03.02.2015 Views

Integral geometry in complex space forms

Integral geometry in complex space forms

Integral geometry in complex space forms

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Integral</strong> <strong>geometry</strong> <strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong><br />

Judit Abardia Bochaca<br />

October 2009<br />

Memòria presentada per aspirar al grau de<br />

Doctor en Ciències Matemàtiques.<br />

Departament de Matemàtiques de la Universitat<br />

Autònoma de Barcelona.<br />

Directors:<br />

Eduardo Gallego Gómez i Gil Solanes Farrés


CERTIFIQUEM que la present Memòria ha estat realitzada per na<br />

Judit Abardia Bochaca, sota la direcció dels Drs. Eduardo Gallego<br />

Gómez i Gil Solanes Farrés.<br />

Bellaterra, octubre del 2009<br />

Signat: Dr. Eduardo Gallego Gómez i Dr. Gil Solanes Farrés


Contents<br />

Introduction 1<br />

1 Spaces of constant holomorphic curvature 7<br />

1.1 First def<strong>in</strong>itions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.2 Projective model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.2.1 Po<strong>in</strong>ts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.2.2 Tangent <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.2.3 Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.2.4 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.2.5 Isometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.2.6 Structure of homogeneous <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . 13<br />

1.3 Mov<strong>in</strong>g frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.4 Submanifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

1.4.1 Totally geodesic submanifolds . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

1.4.2 Geodesic balls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

1.5 Space of <strong>complex</strong> r-planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

1.5.1 Expression for the <strong>in</strong>variant density <strong>in</strong> terms of a parametrization . . . 20<br />

1.5.2 Density of <strong>complex</strong> r-planes conta<strong>in</strong><strong>in</strong>g a fixed <strong>complex</strong> q-plane . . . . . 25<br />

1.5.3 Density of <strong>complex</strong> q-planes conta<strong>in</strong>ed <strong>in</strong> a fixed <strong>complex</strong> r-plane . . . . 25<br />

1.5.4 Measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g a geodesic ball . . . . . . . . . 25<br />

1.5.5 Reproductive property of Quermass<strong>in</strong>tegrale . . . . . . . . . . . . . . . . 26<br />

2 Introduction to valuations 29<br />

2.1 Def<strong>in</strong>ition and basic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.2 Hadwiger Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

2.3 Alesker Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

2.4 Valuations on <strong>complex</strong> <strong>space</strong> <strong>forms</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.4.1 Smooth valuations on manifolds . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.4.2 Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.4.3 Relation between Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes and the valuations given<br />

by Park . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

2.4.4 Other curvature <strong>in</strong>tegrals . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

2.4.5 Relation between the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes and the second fundamental<br />

form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3 Average of the mean curvature <strong>in</strong>tegral 45<br />

3.1 Previous lemmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.2 <strong>Integral</strong> of the r-th mean curvature <strong>in</strong>tegral . . . . . . . . . . . . . . . . . . . . 48<br />

3.3 Mean curvature <strong>in</strong>tegral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

3.4 Reproductive valuations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

v


vi<br />

Contents<br />

3.5 Relation with some valuations def<strong>in</strong>ed by Alesker . . . . . . . . . . . . . . . . . 58<br />

3.6 Example: sphere <strong>in</strong> CK 3 (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

4 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes 61<br />

4.1 Variation of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes . . . . . . . . . . . . . . . . . . . . 61<br />

4.2 Variation of the measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g a doma<strong>in</strong> . . . . . . . 66<br />

4.3 Measure of <strong>complex</strong> r-planes meet<strong>in</strong>g a regular doma<strong>in</strong> . . . . . . . . . . . . . . 68<br />

4.3.1 In the standard Hermitian <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . 68<br />

4.3.2 In <strong>complex</strong> <strong>space</strong> <strong>forms</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.4 Gauss-Bonnet formula <strong>in</strong> CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

4.5 Another method to compute the measure of <strong>complex</strong> l<strong>in</strong>es . . . . . . . . . . . . 76<br />

4.5.1 Measure of <strong>complex</strong> l<strong>in</strong>es meet<strong>in</strong>g a regular doma<strong>in</strong> <strong>in</strong> C n . . . . . . . . 76<br />

4.5.2 Measure of <strong>complex</strong> l<strong>in</strong>es meet<strong>in</strong>g a regular doma<strong>in</strong> <strong>in</strong> CP n and CH n . . 77<br />

4.6 Total Gauss curvature <strong>in</strong>tegral C n . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

5 Other Crofton formulas 81<br />

5.1 Space of (k, p)-planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

5.1.1 Bisectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5.2 Variation of the measure of planes meet<strong>in</strong>g a regular doma<strong>in</strong> . . . . . . . . . . 84<br />

5.3 Measure of real geodesics <strong>in</strong> CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

5.4 Measure of real hyperplanes <strong>in</strong> C n . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

5.5 Measure of coisotropic planes <strong>in</strong> C n . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5.6 Measure of Lagrangian planes <strong>in</strong> CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . 90<br />

Appendix 93<br />

Proof of Theorem 4.3.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

Proof of Theorem 4.4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

Bibliography 103<br />

Notation 107<br />

Index 109


Introduction<br />

Classically, <strong>in</strong>tegral <strong>geometry</strong> <strong>in</strong> Euclidean <strong>space</strong> deals with two basic questions: the expression<br />

of the measure of planes meet<strong>in</strong>g a convex doma<strong>in</strong>, the so-called Crofton formulas; and the<br />

study of the measure of movements tak<strong>in</strong>g one convex doma<strong>in</strong> over another fixed convex<br />

doma<strong>in</strong>, the so-called k<strong>in</strong>ematic formula.<br />

In the Euclidean <strong>space</strong> R n , we denote by L r a totally geodesic submanifold of dimension<br />

r, and we call it r-plane. We denote the <strong>space</strong> of r-planes by L r . This <strong>space</strong> has a unique<br />

(up to a constant factor) density <strong>in</strong>variant under the isometry group of R n , denoted by dL r .<br />

Then, given a convex doma<strong>in</strong> Ω ⊂ R n with smooth boundary, the expression of the measure<br />

of r-planes meet<strong>in</strong>g a convex doma<strong>in</strong> is given by<br />

∫<br />

L r<br />

χ(Ω ∩ L r )dL r = c n,r M r−1 (∂Ω), (1)<br />

where c n,r only depends on the dimensions n, r, and M r−1 (∂Ω) denotes the <strong>in</strong>tegral over ∂Ω<br />

of the (r − 1)-th mean curvature <strong>in</strong>tegral.<br />

Thus, the mean curvature <strong>in</strong>tegrals appear naturally <strong>in</strong> the Crofton formula. A classical<br />

known property of mean curvature <strong>in</strong>tegrals is the follow<strong>in</strong>g<br />

∫<br />

M (r)<br />

i<br />

(∂Ω ∩ L r )dL r = c ′ n,r,iM i (∂Ω) (2)<br />

L r<br />

where c ′ (r)<br />

n,r,i only depends on the dimensions n, r, i, and M<br />

i<br />

(∂Ω ∩ L r ) denotes the i-th mean<br />

curvature <strong>in</strong>tegral of ∂Ω ∩ L r as a hypersurface <strong>in</strong> L r<br />

∼ = R r . From (2), it is said that mean<br />

curvature <strong>in</strong>tegrals satisfy a reproductive property.<br />

On the other hand, the k<strong>in</strong>ematic formula <strong>in</strong> R n is expressed as follows. Let Ω 1 and Ω 2 be<br />

two convex doma<strong>in</strong>s with smooth boundary, let O(n) := O(n) ⋉ R n denote the isometry group<br />

of R n , and let dg be an <strong>in</strong>variant density of O(n). Then,<br />

∫<br />

O(n)<br />

χ(Ω 1 ∩ gΩ 2 )dg =<br />

n∑<br />

c n,i M i (∂Ω 1 )M n−i (∂Ω 2 ). (3)<br />

i=0<br />

The previous three formulas were extended to projective and hyperbolic <strong>space</strong>s (cf. [San04]),<br />

i.e. they are known <strong>in</strong> the <strong>space</strong>s of constant sectional curvature k. The generalization of <strong>in</strong>tegral<br />

(2) does not depend on k but <strong>in</strong> the expression (1) for projective and hyperbolic <strong>space</strong><br />

appear other terms, depend<strong>in</strong>g on k. Moreover, its expression depends on the parity of the<br />

dimension of the planes. If r is even, then<br />

∫<br />

L r<br />

χ(Ω ∩ L r )dL r = c n,r−1 M r−1 (∂Ω) + c n,r−3 M r−3 (∂Ω) + · · · + c n,1 M 1 (∂Ω) + c n vol(Ω), (4)<br />

and if r is odd<br />

∫<br />

L r<br />

χ(Ω ∩ L r )dL r = c n,r−1 M r−1 (∂Ω) + c n,r−3 M r−3 (∂Ω) + · · · + c n,2 M 2 (∂Ω) + c n vol(∂Ω), (5)<br />

1


2 Introduction<br />

where c n,j depends on the dimensions n and j and are multiples of k n−j .<br />

The facts that the expression depends on the parity, and that we study an <strong>in</strong>tegral of<br />

the Euler characteristic, rema<strong>in</strong> us the Gauss-Bonnet formula <strong>in</strong> <strong>space</strong>s of constant sectional<br />

curvature, which also depends on the parity of the ambient <strong>space</strong>. We recall here this formula<br />

<strong>in</strong> a <strong>space</strong> of constant sectional curvature k and dimension n.<br />

If n is even, then<br />

M n−1 (∂Ω) + c n−3 M n−3 (∂Ω) + · · · + c 1 M 1 (∂Ω) + k n/2 vol(Ω) = vol(S n−1 )χ(Ω),<br />

and if n is odd,<br />

M n−1 (∂Ω) + c n−3 M n−3 (∂Ω) + · · · + c 2 M 2 (∂Ω) + k (n−1)/2 vol(∂Ω) = vol(Sn−1 )<br />

χ(Ω)<br />

2<br />

where c i depends only on the dimensions n, i and are multiples of the sectional curvature k.<br />

Now, us<strong>in</strong>g the expression (2) and the Gauss-Bonnet formula, we get (4) and (5).<br />

The goal of this work is generalize formulas (1) and (2) <strong>in</strong> the standard Hermitian <strong>space</strong><br />

C n , <strong>in</strong> the <strong>complex</strong> projective <strong>space</strong> and <strong>in</strong> the <strong>complex</strong> hyperbolic <strong>space</strong>, denoted by CK n (ɛ)<br />

with 4ɛ the holomorphic curvature of the manifold (see Section 1.1).<br />

In order to achieve this goal, we use the notion of valuation <strong>in</strong> a vector <strong>space</strong> V , a realvalued<br />

functional φ from the <strong>space</strong> of convex compact doma<strong>in</strong>s K(V ) <strong>in</strong> V to R satisfy<strong>in</strong>g the<br />

follow<strong>in</strong>g additive property<br />

φ(A ∪ B) = φ(A) + φ(B) − φ(A ∩ B)<br />

whenever A, B, A ∪ B ∈ K(V ).<br />

The first examples of valuations are the volume of the convex doma<strong>in</strong>, the area of the<br />

boundary, and the Euler characteristic. Other classical examples of valuations are the socalled<br />

<strong>in</strong>tr<strong>in</strong>sic volumes. They are def<strong>in</strong>ed from the Ste<strong>in</strong>er formula: given a convex doma<strong>in</strong><br />

Ω ⊂ R n , if we denote by Ω r the parallel doma<strong>in</strong> at a distance r, the Ste<strong>in</strong>er formula relates<br />

the volume of Ω r with the so-called <strong>in</strong>tr<strong>in</strong>sic volumes V i (Ω) by<br />

vol(Ω r ) =<br />

n∑<br />

r n−i ω n−i V i (Ω)<br />

i=0<br />

where ω n−i denotes the volume of the (n − i)-dimensional Euclidean ball with radius 1 (cf.<br />

Proposition 2.1.3).<br />

If Ω ⊂ R n is a convex doma<strong>in</strong> with smooth boundary, then <strong>in</strong>tr<strong>in</strong>sic volumes satisfy<br />

V i (Ω) = cM n−i−1 (∂Ω),<br />

and they are the natural generalization of the mean curvature <strong>in</strong>tegrals for non-smooth convex<br />

doma<strong>in</strong>s.<br />

Hadwiger <strong>in</strong> [Had57] proved that all cont<strong>in</strong>uous valuations <strong>in</strong> R n <strong>in</strong>variant under the isometry<br />

group of R n are l<strong>in</strong>ear comb<strong>in</strong>ation of the volume of the convex doma<strong>in</strong>, the area of<br />

the boundary, and the <strong>in</strong>tr<strong>in</strong>sic volumes (see Section 2.2.1). This result has as immediate<br />

consequence formulas (1), (2) and (3).<br />

Alesker <strong>in</strong> [Ale03] proved that the dimension of the <strong>space</strong> of cont<strong>in</strong>uous valuations <strong>in</strong> C n<br />

<strong>in</strong>variant under the holomorphic isometry group of C n is ( )<br />

n+2<br />

2 and gave a basis of this <strong>space</strong>.<br />

In the recent paper of Bernig and Fu, [BF08], there are given other basis of valuations <strong>in</strong> C n .<br />

In particular, the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes are def<strong>in</strong>ed (see Section 2.4.2). These are the<br />

valuations we will use to work with. The fact that the dimension of the <strong>space</strong> of cont<strong>in</strong>uous<br />

valuations <strong>in</strong>variant under the isometry group of C n is bigger than the one of R 2n is not


3<br />

surpris<strong>in</strong>g if we recall that the holomorphic isometry group of C n , U(n), is smaller than the<br />

isometry group of R 2n , O(2n).<br />

Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes are a k<strong>in</strong>d of generalization of mean curvature <strong>in</strong>tegrals, but<br />

tak<strong>in</strong>g <strong>in</strong>to account that C n has a <strong>complex</strong> structure which def<strong>in</strong>es a canonical vector field on<br />

hypersurfaces. Indeed, at each po<strong>in</strong>t x of a hypersurface, if we consider the normal vector, and<br />

we apply the <strong>complex</strong> structure, then we get a dist<strong>in</strong>guished vector JN <strong>in</strong> the tangent <strong>space</strong><br />

of the hypersurface at x. Moreover, the orthogonal <strong>space</strong> to JN <strong>in</strong> the tangent <strong>space</strong> def<strong>in</strong>es<br />

a <strong>complex</strong> <strong>space</strong> of maximum dimension, n − 1.<br />

So, if S is a smooth hypersurface <strong>in</strong> C n , we can consider the <strong>in</strong>tegral<br />

∫<br />

k n (JN)dx<br />

S<br />

where k n (JN) denotes the normal curvature <strong>in</strong> the direction JN, and this is a valuation <strong>in</strong><br />

C n . Other valuations related to normal curvature of the direction JN appear as elements <strong>in</strong><br />

of Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes basis.<br />

The notion of valuation can be also def<strong>in</strong>ed <strong>in</strong> a differentiable manifold (see Def<strong>in</strong>ition<br />

2.4.1). In real <strong>space</strong> <strong>forms</strong> the volume of a convex doma<strong>in</strong> and the area of its boundary are<br />

valuations. But, it is not known an analogous result to Hadwiger Theorem <strong>in</strong> these <strong>space</strong>s.<br />

The def<strong>in</strong>ition of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes can be extended to other <strong>space</strong> of constant<br />

holomorphic curvature. We denote by {µ k,q } the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes. The subscript<br />

k denotes the degree of the valuation (see Section 2.4.2).<br />

In order to give a similar expression of (1) and (2) <strong>in</strong> the <strong>space</strong>s of constant holomorphic<br />

curvature, we need to describe the <strong>in</strong>tegration <strong>space</strong>. Note that <strong>in</strong> <strong>space</strong>s of constant sectional<br />

curvature we <strong>in</strong>tegrate over the <strong>space</strong> of r-planes, i.e. totally geodesic submanifold of fixed dimension.<br />

In <strong>space</strong>s of constant holomorphic curvature, complete totally geodesic submanifolds<br />

are classified. If ɛ ≠ 0 they are <strong>complex</strong> submanifolds isometric to CK r (ɛ) ⊂ CK n (ɛ), with<br />

1 ≤ r < n or totally real submanifolds isometric to RK q (ɛ) ⊂ CK n (ɛ) with 1 ≤ q ≤ n, where<br />

RK q (ɛ) denotes the <strong>space</strong> of constant sectional curvature ɛ. For ɛ = 0 there are other totally<br />

geodesic submanifolds. We denote the <strong>space</strong> of <strong>complex</strong> planes with <strong>complex</strong> dimension r,<br />

1 ≤ r < n, by L C r , and the <strong>space</strong> of totally real planes of maximum dimension n by L R n, the<br />

so-called Lagrangian manifolds.<br />

In this work, we obta<strong>in</strong> a Crofton formula for <strong>complex</strong> r-planes and Lagrangian planes<br />

∫<br />

( ) n − 1<br />

χ(Ω ∩ L r )dL r = vol(G C n−1,r)<br />

−1· (6)<br />

L C r<br />

· (<br />

n−1<br />

∑<br />

k=n−r<br />

ɛ k−(n−r) ω 2n−2k<br />

( n<br />

k<br />

+ ɛ r (r + 1)vol(Ω)),<br />

) −1<br />

⎛<br />

⎝<br />

r<br />

∑k−1<br />

q=max{0,2k−n}<br />

( 2k−2q<br />

)<br />

⎞<br />

k−q<br />

4 k−q µ 2k,q (Ω) + (k + r − n + 1)µ 2k,k (Ω) ⎠<br />

and<br />

∫<br />

L R n<br />

⎛<br />

· ⎝<br />

∫<br />

L R n<br />

χ(Ω ∩ L)dL = vol(G 2n,n)ω n<br />

n!<br />

n−1<br />

2∑<br />

q=0<br />

( ) 2q − 1 −1<br />

4 q−n<br />

q − 1 2q + 1 µ n,q(Ω) if n is odd, (7)<br />

χ(Ω ∩ L)dL = vol(G 2n,n)<br />

· (8)<br />

n!<br />

n<br />

2∑<br />

( ) 2q − 1 −1<br />

4 q−n n<br />

⎞<br />

ω n<br />

2∑<br />

( ) n −1<br />

q − 1 2q + 1 µ n,q(Ω) + ɛ i 2 −n+1 ω n−2i<br />

n<br />

2 + i µ<br />

n + 1 n+2i,<br />

n<br />

2 +i (Ω) ⎠ if n is even,<br />

q=0<br />

i=1


4 Introduction<br />

where ω i denotes the volume of the i-dimensional Euclidean unit ball.<br />

Previous formulas have more addends that the correspond<strong>in</strong>g ones <strong>in</strong> the <strong>space</strong>s of constant<br />

sectional curvature, but they are similar. If ɛ = 0, i.e. <strong>in</strong> C n also appear all the valuations<br />

with the correspond<strong>in</strong>g degree. If ɛ ≠ 0 the notion of degree of a valuation has no sense but<br />

there is a similitude with the expression <strong>in</strong> <strong>space</strong>s of constant sectional curvature compar<strong>in</strong>g<br />

the subscripts of the valuations.<br />

In order to get these expressions we use a variational method. That is, we take a smooth<br />

vector field X def<strong>in</strong>ed on the manifold and we consider its flow φ t . We prove the follow<strong>in</strong>g<br />

formula of first variation<br />

∫<br />

∫ ∫<br />

d<br />

dt∣ χ(φ t (Ω) ∩ L r )dL r = 〈X, N〉<br />

σ 2r (II| V )dV dp<br />

t=0 G C n−1,r (Dp)<br />

L C r<br />

∂Ω<br />

where N is the exterior normal field, D is the distribution <strong>in</strong> the tangent <strong>space</strong> at ∂Ω orthogonal<br />

to JN, and σ 2r (II| V ) denotes the 2r-th symmetric elementary function of the second<br />

fundamental form II restricted to V ∈ G C n−1,r (D p), the Grassmannian of <strong>complex</strong> planes with<br />

<strong>complex</strong> dimension r <strong>in</strong>side D p .<br />

On the other hand, we get an expression of the variation of valuations µ k,q us<strong>in</strong>g the method<br />

<strong>in</strong> [BF08].<br />

Compar<strong>in</strong>g both variations and solv<strong>in</strong>g a system of l<strong>in</strong>ear equations we obta<strong>in</strong> the result.<br />

Us<strong>in</strong>g the same variational method we also obta<strong>in</strong> a Gauss-Bonnet formula for the <strong>space</strong>s<br />

of constant holomorphic curvature. It is known that the variation of the Euler characteristic<br />

is zero. Thus, we can express it as a sum of Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes such that its variation<br />

vanishes. The obta<strong>in</strong>ed Gauss-Bonnet formula is the follow<strong>in</strong>g<br />

⎛<br />

n−1<br />

∑<br />

ω 2n χ(Ω) = (n + 1)ɛ n ɛ c ω 2n−2c (n − c)<br />

vol(Ω) +<br />

n ( ) ⎝<br />

n−1<br />

c<br />

c=0<br />

∑c−1<br />

q=max{0,2c−n}<br />

( 2c−2q<br />

)<br />

⎞<br />

c−q<br />

4 c−q µ 2c,q + (c + 1)µ 2c,c<br />

⎠.<br />

In <strong>space</strong>s of constant sectional curvature k, Solanes <strong>in</strong> [Sol06] related the measure of planes<br />

meet<strong>in</strong>g a doma<strong>in</strong> with the Euler characteristic of the doma<strong>in</strong><br />

In CK n (ɛ), we get<br />

ω n χ(Ω) = 1 n M n−1(∂Ω) +<br />

2k ∫<br />

nω n−1<br />

L n−2<br />

χ(Ω ∩ L n−2 )dL n−2 .<br />

ω 2n χ(Ω) = 1<br />

∫<br />

2n M 2n−1(∂Ω) + ɛ χ(Ω ∩ L n−1 )dL n−1 +<br />

L C n−1<br />

n∑<br />

j=1<br />

ɛ j ω 2n<br />

µ 2j,j (Ω).<br />

ω 2j<br />

(9)<br />

The analogous expression to (2), it is given when we <strong>in</strong>tegrate the mean curvature <strong>in</strong>tegral<br />

over <strong>complex</strong> r-planes. The obta<strong>in</strong>ed expression for a compact oriented (possible with<br />

boundary) hypersurface S of class C 2 is<br />

∫<br />

L C r<br />

M (r)<br />

1 (S∩L r)dL r = ω 2n−2vol(G C n−2,r−1 ) ( n −1 (<br />

(2n − 1)<br />

2r(2r − 1) r) 2nr − n − r ∫<br />

M 1 (S) +<br />

n − r<br />

where k n (JN) denotes the normal curvature <strong>in</strong> the direction JN ∈ T S.<br />

S<br />

)<br />

k n (JN)<br />

(10)


5<br />

To get this result, first we obta<strong>in</strong>, us<strong>in</strong>g mov<strong>in</strong>g frames, the follow<strong>in</strong>g <strong>in</strong>termediate expression<br />

(for any mean curvature <strong>in</strong>tegral). If r, i ∈ N such that 1 ≤ r ≤ n and 0 ≤ i ≤ 2r − 1,<br />

then<br />

∫<br />

M (r)<br />

i<br />

(S ∩ L r )dL r (11)<br />

L C r<br />

( ) 2r − 1 −1 ∫<br />

=<br />

i<br />

S<br />

∫RP 2n−2 ∫<br />

G C n−2,r−1<br />

|〈JN, e r 〉| 2r−i<br />

(1 − 〈JN, e r 〉 2 ) r−1 σ i(p; e r ⊕ V )dV de r dp,<br />

where e r ∈ T p S unit vector, V denotes a <strong>complex</strong> (r − 1)-plane conta<strong>in</strong><strong>in</strong>g p and conta<strong>in</strong>ed <strong>in</strong><br />

{N, JN, e r , Je r } ⊥ , σ i (p; e r ⊕ V ) denotes the i-th symmetric elementary function of the second<br />

fundamental form of S restricted to the real sub<strong>space</strong> e r ⊕ V and the <strong>in</strong>tegration over RP 2n−2<br />

denotes the projective <strong>space</strong> of the unit tangent <strong>space</strong> of the hypersurface.<br />

In order to complete the generalization of equation (1) <strong>in</strong> CK n (ɛ), it rema<strong>in</strong>s to study the<br />

measure of (non-maximal) totally real planes. These are the other totally geodesic submanifolds<br />

of CK n (ɛ), ɛ ≠ 0. Us<strong>in</strong>g the same techniques as <strong>in</strong> the rest of this work, it does not<br />

seem possible to solve this case s<strong>in</strong>ce we cannot obta<strong>in</strong> enough <strong>in</strong>formation of the variational<br />

properties of the measures of totally real planes meet<strong>in</strong>g a doma<strong>in</strong> <strong>in</strong> CK n (ɛ).<br />

On the other hand, it would be <strong>in</strong>terest<strong>in</strong>g to extend formula (10) to i ∈ {2, . . . , 2r − 1}.<br />

Next we expla<strong>in</strong> the organization of the text.<br />

Chapter 1 conta<strong>in</strong>s a description of the <strong>space</strong>s of constant holomorphic curvature. We<br />

review its def<strong>in</strong>ition and describe some of the most important submanifolds, i.e. the totally<br />

geodesic submanifolds, the geodesic spheres, and the <strong>complex</strong> planes. In this chapter we also<br />

recall the method of mov<strong>in</strong>g frames, which will be used along this text. Us<strong>in</strong>g mov<strong>in</strong>g frames,<br />

we give an expression for the density of the <strong>space</strong> of <strong>complex</strong> planes. F<strong>in</strong>ally, we prove that<br />

<strong>in</strong>tegral ∫ L χ(Ω ∩ L r)dL C r satisfies a reproductive property.<br />

r<br />

Chapter 2 is devoted to the study of valuations <strong>in</strong> C n and <strong>in</strong> the <strong>space</strong>s of constant holomorphic<br />

curvature. First of all, we review the concept and the ma<strong>in</strong> properties of the valuations<br />

on R n together with the Hadwiger Theorem (which characterize all cont<strong>in</strong>uous valuations <strong>in</strong><br />

R n <strong>in</strong>variants under the isometry group). An analogous Hadwiger Theorem <strong>in</strong> C n is stated.<br />

F<strong>in</strong>ally, we def<strong>in</strong>e the used valuations <strong>in</strong> this work <strong>in</strong> <strong>space</strong>s of constant holomorphic curvature,<br />

and we give new properties and relations with other valuations also important <strong>in</strong> the<br />

next chapters.<br />

Chapter 3 gives a proof of (10). First of all, we prove some geometric lemmas and we<br />

obta<strong>in</strong> the expression for the mean curvature <strong>in</strong>tegrals over the <strong>space</strong> of <strong>complex</strong> planes <strong>in</strong><br />

terms of an <strong>in</strong>tegral over the boundary of the doma<strong>in</strong> given <strong>in</strong> (11). This expression will be<br />

fundamental to atta<strong>in</strong> the goal of this chapter. As a corollary of (10) we characterize the<br />

valuations of degree 2n − 2 satisfy<strong>in</strong>g a reproductive property <strong>in</strong> C n , and we give the relation<br />

among different valuations def<strong>in</strong>ed by Alesker (already reviewed <strong>in</strong> Chapter 2). The results of<br />

this chapter are conta<strong>in</strong>ed <strong>in</strong> [Aba].<br />

In Chapter 4 we obta<strong>in</strong> the measure of <strong>complex</strong> planes <strong>in</strong>tersect<strong>in</strong>g a doma<strong>in</strong> <strong>in</strong> the <strong>space</strong>s of<br />

constant holomorphic curvature <strong>in</strong> terms of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes def<strong>in</strong>ed at Chapter<br />

2. We also give an expression of the Gauss-Bonnet formula <strong>in</strong> terms of these valuations. In<br />

order to get these expressions we use a variational method. First, we obta<strong>in</strong> an expression for<br />

the variation of the measure of <strong>complex</strong> planes and for the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes. In<br />

this chapter, we verify the certa<strong>in</strong>ty of (6). A constructive proof, where we f<strong>in</strong>d the constants<br />

<strong>in</strong> the expression is given <strong>in</strong> the appendix. As a corollary, we express the total mean Gauss<br />

curvature <strong>in</strong> C n also <strong>in</strong> terms of the Hermitic <strong>in</strong>tr<strong>in</strong>sic volumes. F<strong>in</strong>ally, we relate Chapters 3


6 Introduction<br />

and 4 obta<strong>in</strong><strong>in</strong>g another method to compute the measure of <strong>complex</strong> l<strong>in</strong>es meet<strong>in</strong>g a doma<strong>in</strong>.<br />

The results of this chapter are conta<strong>in</strong>ed <strong>in</strong> [AGS09].<br />

Chapter 5 studies the measure of another type of planes meet<strong>in</strong>g a doma<strong>in</strong> <strong>in</strong> C n , the socalled<br />

coisotropic planes. These planes are the orthogonal direct sum of a <strong>complex</strong> sub<strong>space</strong><br />

of <strong>complex</strong> dimension n − p and a totally real sub<strong>space</strong> of dimension p. Totally real planes of<br />

maximum dimension and real hyperplanes are particular cases of this type of planes. Us<strong>in</strong>g<br />

similar techniques as <strong>in</strong> Chapter 4 we give an expression for the measure of planes of this type<br />

meet<strong>in</strong>g a doma<strong>in</strong>. For the <strong>space</strong>s of constant holomorphic curvature we prove (7) and (8),<br />

which give the measure of totally real planes of maximum dimension, the so-called Lagrangian<br />

planes.<br />

The appendix conta<strong>in</strong>s the constructive proof of (6) and (9). That is, we give the method<br />

that allowed us to obta<strong>in</strong> the constants appear<strong>in</strong>g <strong>in</strong> these expressions. This proof consists,<br />

at a f<strong>in</strong>al <strong>in</strong>stance, to solve a l<strong>in</strong>ear system obta<strong>in</strong>ed from the study of the variation of both<br />

sides of the expressions, as it is detailed <strong>in</strong> Chapter 4.<br />

Acknowledgments<br />

In these f<strong>in</strong>al l<strong>in</strong>es, I would like to express my gratitude to all people who, <strong>in</strong> some way, allowed<br />

me to convert this work <strong>in</strong>to a Thesis. Specially, I would like to mention my first director,<br />

Eduardo Gallego, my codirector, Gil Solanes, for their patient and support; and the Group<br />

of Geometry and Topology of the Universitat Autònoma de Barcelona, <strong>in</strong> particular Agustí<br />

Reventós, to <strong>in</strong>troduce me to the “<strong>geometry</strong> world”. I would also like to thank Andreas Bernig<br />

for all his advices, and the members of the Department of Mathematics <strong>in</strong> the University of<br />

Fribourg to offer me such nice work ambience dur<strong>in</strong>g my stay. Although I am not go<strong>in</strong>g to<br />

write down a list with the name of all mathematicians who contributed with their enlighten<strong>in</strong>g<br />

conversations to the completion of this work, and hopefully, to some future work, I would be<br />

very satisfied if they feel identified <strong>in</strong> these l<strong>in</strong>es.<br />

Outside from the direct relation to this work, I would like to thank my colleagues of<br />

doctorate, specially my colleagues <strong>in</strong> the room, and also <strong>in</strong> the <strong>geometry</strong> and topology area,<br />

<strong>in</strong> Barcelona, but also <strong>in</strong> Fribourg. In the same way I would like to thank all the persons I<br />

have met dur<strong>in</strong>g these years and become my friends.<br />

F<strong>in</strong>ally, I thank my parent’s support, not only <strong>in</strong> these years of doctorate.<br />

This work has been written under the support of the “Departament d’Universitats, Recerca<br />

i Societat de la Informació de la Generalitat de Catalunya”, the European Social Found and<br />

the Swiss National Found.


Chapter 1<br />

Spaces of constant holomorphic<br />

curvature<br />

1.1 First def<strong>in</strong>itions<br />

In this section we <strong>in</strong>troduce the <strong>space</strong>s of constant holomorphic curvature, also called <strong>complex</strong><br />

<strong>space</strong> <strong>forms</strong>, and give the properties we shall use along this work. First of all, we recall some<br />

basic def<strong>in</strong>itions.<br />

Def<strong>in</strong>ition 1.1.1. Let M be a differentiable manifold. M is a <strong>complex</strong> manifold if it has an<br />

atlas such that the change of coord<strong>in</strong>ates are holomorphic, that is, {(U α , φ α )} is an atlas with<br />

{U α } an open cover<strong>in</strong>g of M and φ α : U α → C n homeomorphisms such that φ β ◦ φ −1<br />

α are<br />

holomorphic <strong>in</strong> its doma<strong>in</strong> of def<strong>in</strong>ition.<br />

Examples<br />

(i) The vector <strong>space</strong> C n is a <strong>complex</strong> manifold of <strong>complex</strong> dimension n.<br />

(ii) The <strong>complex</strong> projective <strong>space</strong> CP n is a <strong>complex</strong> manifold of <strong>complex</strong> dimension n.<br />

The <strong>complex</strong> projective <strong>space</strong> can be def<strong>in</strong>ed analogously to the real projective <strong>space</strong>.<br />

Let us consider <strong>in</strong> C n+1 \{0} the equivalence relation which identifies the po<strong>in</strong>ts differ<strong>in</strong>g<br />

by a <strong>complex</strong> multiple. Then, we take as an atlas the open sets {U 0 , ..., U n } such that<br />

U j = {(z 0 , ..., z n ) ∈ C n+1 | z j ≠ 0}<br />

and for every U j we take the map φ j (z 0 , ..., z n ) = (z 0 /z j , ..., z j−1 /z j , z j+1 /z j , ..., z n /z j )<br />

which is a homeomorphism. It can be proved that the change of coord<strong>in</strong>ates are holomorphics.<br />

Def<strong>in</strong>ition 1.1.2. Let M be a <strong>complex</strong> manifold. A l<strong>in</strong>ear map J : T x M → T x M is an almost<br />

<strong>complex</strong> structure of M if for each x ∈ M, the restriction of J at T x M satisfies J x : T x M →<br />

T x M, J 2 = −Id, and J varies differentially on M.<br />

Note that any <strong>complex</strong> manifold admits an almost <strong>complex</strong> structure. Indeed, the tangent<br />

<strong>space</strong> of a <strong>complex</strong> manifold has a <strong>complex</strong> vector <strong>space</strong> structure, so the map “multiply by i”<br />

is well-def<strong>in</strong>ed and satisfies that applied twice is the map −Id. We call this canonical almost<br />

<strong>complex</strong> structure <strong>complex</strong> structure.<br />

Def<strong>in</strong>ition 1.1.3. Let V be a <strong>complex</strong> vector <strong>space</strong> and let u, v ∈ V . It is said that h :<br />

V × V → C is an Hermitian product on V if<br />

7


8 Spaces of constant holomorphic curvature<br />

1. it is C-l<strong>in</strong>ear with respect to the first component,<br />

2. h(u, v) = h(u, v).<br />

Remark 1.1.4. From the properties of a Hermitian product, it follows that if λ ∈ C then<br />

h(u, v) = λh(u, v). Indeed, by def<strong>in</strong>ition we get the follow<strong>in</strong>g equalities<br />

h(u, λv) = h(λv, u) = λh(v, u) = λh(u, v).<br />

Def<strong>in</strong>ition 1.1.5. Let M be a differentiable manifold with <strong>complex</strong> structure J and a Riemannian<br />

metric g. Then, g is called a Hermitian metric if it is compatible with the <strong>complex</strong><br />

structure, i.e. it satisfies g x (Ju, Jv) = g x (u, v) for every x ∈ M and u, v ∈ T x M.<br />

Def<strong>in</strong>ition 1.1.6. Let M be a <strong>complex</strong> manifold with <strong>complex</strong> structure J and Hermitian<br />

metric g. The 2-form ω def<strong>in</strong>ed by<br />

is the Kähler form.<br />

ω(u, v) = g(u, Jv), ∀ u, v ∈ T x M<br />

Remark 1.1.7. Given a <strong>complex</strong> manifold M with <strong>complex</strong> structure J and a Hermitian product<br />

def<strong>in</strong>ed on T M, we get a Hermitian metric on M from the real part of the Hermitian product,<br />

and a Kähler form on M from the imag<strong>in</strong>ary part of the Hermitian product.<br />

Def<strong>in</strong>ition 1.1.8. A <strong>complex</strong> manifold M is called a Kähler manifold if it has a Hermitian<br />

metric such that the Kähler form associated to this metric is closed.<br />

Proposition 1.1.9 ([O’N83] page 326). Let M be a Kähler manifold with connection ∇.<br />

Then,<br />

∇JX = J∇X, ∀X ∈ X(M).<br />

Def<strong>in</strong>ition 1.1.10. A sub<strong>space</strong> W ⊂ T x M of <strong>complex</strong> dimension 1 is a <strong>complex</strong> direction or<br />

a holomorphic section of the tangent <strong>space</strong> if W is <strong>in</strong>variant under J, i.e. JW = W .<br />

If w ≠ 0 ∈ W , then the vectors {w, Jw} constitute a basis of W , as a real sub<strong>space</strong>.<br />

Def<strong>in</strong>ition 1.1.11. The holomorphic curvature is the sectional curvature of holomorphic sections.<br />

Def<strong>in</strong>ition 1.1.12. A <strong>space</strong> of constant holomorphic curvature 4ɛ of dimension n is a complete,<br />

simply connected Kähler manifold of <strong>complex</strong> dimension n, such that the holomorphic<br />

curvature is constant and equal to 4ɛ for every po<strong>in</strong>t and every <strong>complex</strong> direction.<br />

Theorem 1.1.13 ([KN69] Theorem 7.9 page 170). Two complete, simply connected Kähler<br />

manifolds with constant holomorphic curvature equal to 4ɛ are holomorphically isometric.<br />

Def<strong>in</strong>ition 1.1.14. We denote by CK n (ɛ) any <strong>space</strong> of constant holomorphic curvature of<br />

dimension n. If ɛ > 0, then it corresponds to the <strong>complex</strong> projective <strong>space</strong> CP n , if ɛ < 0, to<br />

the <strong>complex</strong> hyperbolic <strong>space</strong> CH n , and if ɛ = 0, to the Hermitian standard <strong>space</strong> C n .<br />

Spaces of constant holomorphic curvature are also called <strong>complex</strong> <strong>space</strong> <strong>forms</strong>.<br />

Remark 1.1.15. Complex <strong>space</strong> <strong>forms</strong> are, <strong>in</strong> some sense, a generalization of real <strong>space</strong> <strong>forms</strong><br />

(<strong>space</strong>s of constant sectional curvature), i.e. the complete simply connected Riemannian manifolds<br />

with constant sectional curvature. Real <strong>space</strong> <strong>forms</strong> are (up to isometry) the Euclidean<br />

<strong>space</strong> R n , the real projective <strong>space</strong> RP n and the real hyperbolic <strong>space</strong> H n . The results <strong>in</strong> this<br />

work extend some of the classical results <strong>in</strong> <strong>in</strong>tegral <strong>geometry</strong> from real <strong>space</strong> <strong>forms</strong> to <strong>complex</strong><br />

<strong>space</strong> <strong>forms</strong>. Santaló [San52] and Griffiths [Gri78], among others, obta<strong>in</strong>ed some results<br />

of classical <strong>in</strong>tegral <strong>geometry</strong> <strong>in</strong> the standard Hermitian <strong>space</strong> and <strong>in</strong> the <strong>complex</strong> projective<br />

<strong>space</strong> tak<strong>in</strong>g <strong>complex</strong> submanifolds. In this work, we deal with non-empty doma<strong>in</strong>s and, thus,<br />

with real hypersurfaces.


1.2 Projective model 9<br />

Def<strong>in</strong>ition 1.1.16. Given two planes Π and Π ′ of real dimension 2 <strong>in</strong> a vector <strong>space</strong> with a<br />

scalar product, the angle between the two planes is def<strong>in</strong>ed as the <strong>in</strong>fimum among the angles<br />

between a pair of vectors, one <strong>in</strong> Π and the other one <strong>in</strong> Π ′ .<br />

Def<strong>in</strong>ition 1.1.17. Let Π be a plane with real dimension 2 <strong>in</strong> the tangent <strong>space</strong> of a po<strong>in</strong>t <strong>in</strong> a<br />

Kähler manifold with <strong>complex</strong> structure J. The holomorphic angle µ(Π) is the angle between<br />

Π and J(Π).<br />

Proposition 1.1.18 ([KN69] page 167). Let M be a Kähler manifold with <strong>complex</strong> structure<br />

J and Hermitian metric g. The holomorphic angle of a plane Π ⊂ T x M, x ∈ M, is given by<br />

where u, v form an orthonormal basis of Π.<br />

cos µ(Π) = |g(u, Jv)|<br />

Remark 1.1.19. The holomorphic angle of a plane takes values between 0 and π/2. In the<br />

extreme cases we have holomorphic planes, when the holomorphic angle is 0; and totally real<br />

planes def<strong>in</strong>ed as the planes with holomorphic angle π/2.<br />

In a <strong>complex</strong> <strong>space</strong> form, the sectional curvature of any plane can be computed from the<br />

holomorphic curvature and the holomorphic angle of the plane.<br />

Proposition 1.1.20 ([KN69] page 167). Let M be a Kähler manifold with constant holomorphic<br />

curvature 4ɛ. Then, the sectional curvature of any plane Π ⊂ T x M, x ∈ M is given<br />

by<br />

where µ(Π) is the holomorphic angle of the plane Π.<br />

K(Π) = ɛ ( 1 + 3 cos 2 µ(Π) ) (1.1)<br />

Corollary 1.1.21. Sectional curvature of any plane <strong>in</strong> the tangent <strong>space</strong> of a po<strong>in</strong>t <strong>in</strong> a<br />

<strong>complex</strong> <strong>space</strong> form with constant holomorphic curvature 4ɛ lies <strong>in</strong> the <strong>in</strong>terval [ɛ, 4ɛ], if ɛ > 0<br />

and <strong>in</strong> the <strong>in</strong>terval [4ɛ, ɛ], if ɛ < 0.<br />

1.2 Projective model<br />

Along this work, we shall use the projective model of CK n (ɛ), which we describe here briefly<br />

(cf. [Gol99]).<br />

If ɛ = 0, we are consider<strong>in</strong>g the standard Hermitian <strong>space</strong> C n with the standard Hermitian<br />

product. Along this section we suppose ɛ ≠ 0, unless otherwise stated.<br />

1.2.1 Po<strong>in</strong>ts<br />

Endow C n+1 with the Hermitian product<br />

Def<strong>in</strong>e<br />

(z, w) = sign(ɛ)z 0 w 0 +<br />

n∑<br />

z j w j . (1.2)<br />

j=1<br />

H := {z ∈ C n+1 | (z, z) = ɛ}.<br />

H is a real hypersurface of C n+1 (i.e. it has real dimension 2n + 1). We def<strong>in</strong>e the po<strong>in</strong>ts of<br />

CK n (ɛ) as<br />

CK n (ɛ) := π(H)<br />

where<br />

π : C n+1 \{0} → C n+1 \{0}/C ∗ = CP n . (1.3)


10 Spaces of constant holomorphic curvature<br />

Remarks 1.2.1. (i) The fiber of Π for the po<strong>in</strong>ts <strong>in</strong> π(H) =: CK n (ɛ) is S 1 . Indeed, let z,<br />

w ∈ H such that π(z) = π(w). By def<strong>in</strong>ition of π, we have w = αz. On the other hand,<br />

it holds (z, z) = (αz, αz) = ɛ. Thus, α = e iθ with θ ∈ R and π −1 ([z]) ∼ = S 1 .<br />

(ii) If ɛ > 0, then CK n (ɛ) co<strong>in</strong>cides as a subset with CP n . But, if ɛ < 0, then CK n (ɛ) is an<br />

open set of CP n .<br />

The differentiable structure and the structure of a <strong>complex</strong> manifold we take <strong>in</strong> CK n (ɛ) is<br />

the same as the one of an open set of CP n .<br />

1.2.2 Tangent <strong>space</strong><br />

The tangent <strong>space</strong> of a po<strong>in</strong>t z ∈ H is<br />

T z H = {w ∈ C n+1 | Re(z, w) = 0}.<br />

The elements <strong>in</strong> the tangent <strong>space</strong> of π(z) ∈ CK n (ɛ) are obta<strong>in</strong>ed from the image of the<br />

elements <strong>in</strong> the tangent <strong>space</strong> of the po<strong>in</strong>t z under dπ. Moreover, the kernel of dπ has dimension<br />

1.<br />

The direction that its image under dπ is the null vector is Jz, s<strong>in</strong>ce it is the tangent<br />

direction to the fiber. (Note that Jz ∈ T z H s<strong>in</strong>ce Re(z, Jz) = 0.) Indeed, the fiber of a po<strong>in</strong>t<br />

[z] is {e iθ z | θ ∈ R}, then<br />

∂(e iθ z)<br />

∂θ ∣ = iz = Jz.<br />

θ=0<br />

Thus, the tangent <strong>space</strong> at z ∈ H can be decomposed as<br />

T z H = 〈Jz〉 ⊕ 〈Jz〉 ⊥ .<br />

The tangent <strong>space</strong> at po<strong>in</strong>ts <strong>in</strong> CK n (ɛ) co<strong>in</strong>cides with the image by the differential map of<br />

the projection of vectors 〈Jz〉 ⊥ at T z H.<br />

Given a vector v ∈ T π(z) CK n (ɛ) there are <strong>in</strong>f<strong>in</strong>itely many vectors of T z H such that under<br />

the differential map dπ give the same vector v, but we can dist<strong>in</strong>guish the one ly<strong>in</strong>g <strong>in</strong> 〈Jz〉 ⊥ ,<br />

which is called horizontal lift and we denote by v L . All other vectors are obta<strong>in</strong>ed as l<strong>in</strong>ear<br />

comb<strong>in</strong>ation of this vector and a multiple of Jz.<br />

1.2.3 Metric<br />

Let v, w ∈ T π(z) CH n . The Hermitian product at CK n (ɛ) between v, w is def<strong>in</strong>ed by<br />

(v, w) ɛ := (v L , w L ), (1.4)<br />

that is, the Hermitian product def<strong>in</strong>ed at C n+1 applied to the horizontal lift of the vectors.<br />

The real part of this product gives a Hermitian metric on CK n (ɛ)<br />

〈v, w〉 ɛ := Re(v L , w L ).<br />

This metric co<strong>in</strong>cides with the so-called Fub<strong>in</strong>i-Study metric, if ɛ > 0 and with the so-called<br />

Bergmann metric, if ɛ < 0 (cf. [Gol99, page 74]).<br />

Along this work we denote the Hermitian metric of CK n (ɛ) by 〈 , 〉 <strong>in</strong>stead of 〈 , 〉 ɛ .<br />

Notation 1.2.2. In order to unify the study of the <strong>complex</strong> <strong>space</strong> <strong>forms</strong>, we def<strong>in</strong>e, <strong>in</strong> the<br />

same way as it is classically done <strong>in</strong> real <strong>space</strong> <strong>forms</strong>, the follow<strong>in</strong>g trigonometric generalized<br />

functions<br />

⎧<br />

s<strong>in</strong>(α √ ɛ)<br />

√ , if ɛ > 0<br />

⎪⎨ ɛ<br />

s<strong>in</strong> ɛ (α) = α, if ɛ = 0<br />

s<strong>in</strong>h(α √ −ɛ)<br />

⎪⎩ √ , if ɛ < 0<br />

−ɛ


1.2 Projective model 11<br />

and<br />

⎧<br />

⎨ cos(α √ ɛ), if ɛ > 0<br />

cos ɛ (α) = 1, if ɛ = 0<br />

⎩<br />

cosh(α √ |ɛ|), if ɛ < 0<br />

cot ɛ (α) = cos ɛ(α)<br />

s<strong>in</strong> ɛ (α) .<br />

1.2.4 Geodesics<br />

Geodesics <strong>in</strong> the projective model of CK n (ɛ) are given by the projection of the <strong>in</strong>tersection<br />

po<strong>in</strong>ts between H and a plane <strong>in</strong> C n+1 such that it is spanned by a vector correspond<strong>in</strong>g to a<br />

representative <strong>in</strong> H ⊂ C n+1 of a po<strong>in</strong>t z <strong>in</strong> the geodesic at CK n (ɛ), and a vector u tangent to<br />

the geodesic at z.<br />

Then, the expression of a geodesic at CK n (ɛ) is given by [γ(t)] = [cos ɛ (t)z + s<strong>in</strong> ɛ (t)u] where<br />

u ∈ 〈Jz〉 ⊥ ⊂ T z H.<br />

The distance between two po<strong>in</strong>ts <strong>in</strong> the <strong>complex</strong> projective and hyperbolic <strong>space</strong> can be<br />

expressed <strong>in</strong> terms of the Hermitian product def<strong>in</strong>ed at C n+1 .<br />

Proposition 1.2.3 ([Gol99] page 76). Let x, y ∈ CK n (ɛ), ɛ ≠ 0, and let d be the distance<br />

between the two given po<strong>in</strong>ts. If x ′ and y ′ are representatives of x and y, respectively, <strong>in</strong> the<br />

projective model, then the distance between the two po<strong>in</strong>ts is given by<br />

(cos ɛ d(x, y)) 2 = (x′ , y ′ )(y ′ , x ′ )<br />

(x ′ , x ′ )(y ′ , y ′ )<br />

where ( , ) denotes the Hermitian product <strong>in</strong> C n+1 def<strong>in</strong>ed at (1.2).<br />

1.2.5 Isometries<br />

Let us recall the def<strong>in</strong>ition of the matrix Lie group U(p, q).<br />

Def<strong>in</strong>ition 1.2.4. Let (x, y) = − ∑ p−1<br />

j=0 x jy j + ∑ n<br />

j=p x jy j be a Hermitian product <strong>in</strong> C n and<br />

p, q ∈ N ∪ {0} such that p + q = n + 1. Then it is def<strong>in</strong>ed<br />

U(p, q) = {A ∈ M n×n (C) | (Av, Aw) = (v, w) with v, w ∈ C n }.<br />

The matrix group U(n) co<strong>in</strong>cides with U(0, n), that is, we consider the standard Hermitian<br />

product on C n .<br />

The matrices of P U(n + 1) = U(n + 1)/(multiplication by scalars), if ɛ > 0 (resp. the<br />

matrices of P U(1, n) = U(1, n)/(multiplication by <strong>complex</strong> scalars), if ɛ < 0) act naturally on<br />

CK n (ɛ). Moreover, they preserve the metric def<strong>in</strong>ed <strong>in</strong> the model s<strong>in</strong>ce preserve the Hermitian<br />

product def<strong>in</strong>ed at C n+1 . Then, the matrices <strong>in</strong> P U(n + 1) (resp. P U(1, n)) are isometries of<br />

CP n (resp. CH n ).<br />

Proposition 1.2.5 ([Gol99] page 68).<br />

map <strong>in</strong> C n+1 .<br />

• Every isometry of CK n (ɛ) comes from a l<strong>in</strong>ear<br />

• The isometry group of CK n (ɛ) is P U ɛ (n) with<br />

⎧<br />

⎨ C n ⋊ U(n), if ɛ = 0,<br />

U ɛ (n) = U(n + 1) = U(0, n + 1), if ɛ > 0,<br />

⎩<br />

U(1, n), if ɛ < 0.<br />

(1.5)


12 Spaces of constant holomorphic curvature<br />

In order to unify the study of the <strong>complex</strong> <strong>space</strong> <strong>forms</strong>, <strong>in</strong>dependently of ɛ, we represent<br />

the elements <strong>in</strong> the group C n ⋊ U(n) as matrices<br />

( )<br />

1 0<br />

C n . (1.6)<br />

U(n)<br />

The transitivity of the isometry group at different levels is given <strong>in</strong> the follow<strong>in</strong>g proposition.<br />

Proposition 1.2.6 ([Gol99] page 70). The isometry group of CK n (ɛ) acts transitively<br />

• on the po<strong>in</strong>ts <strong>in</strong> CK n (ɛ),<br />

• on the unit tangent bundle. That is, given (p, v), (q, w) <strong>in</strong> the unit tangent bundle there<br />

exists an isometry σ such that σ(p) = q and dσ(v) = w.<br />

• on the holomorphic sections (see Def<strong>in</strong>ition 1.1.10).<br />

In the follow<strong>in</strong>g lemma, we give a basis of left-<strong>in</strong>variant <strong>forms</strong> of U(p, q). We will prove<br />

that these <strong>forms</strong> are also right-<strong>in</strong>variants.<br />

Lemma 1.2.7. Let A = (a 0 , . . . , a m ) ∈ U(p, q), with p, q, m ∈ N∪{0} such that p+q = m+1.<br />

A basis of left-<strong>in</strong>variant <strong>forms</strong> <strong>in</strong> U(p, q) is given by {Re(ϕ jk ), Im(ϕ jk ), Re(ϕ jj )}, 0 ≤ j ≤ k ≤<br />

m, j ≠ k where ϕ ij = (da i , a j ) and (x, y) = − ∑ p−1<br />

j=0 x jy j + ∑ m<br />

j=p x jy j <strong>in</strong> C m+1 .<br />

Proof. From Def<strong>in</strong>ition 1.2.4 of U(p, q) it follows that A ∈ U(p, q) if and only if A −1 = εA t ε<br />

where<br />

( )<br />

−Idp 0<br />

ε =<br />

. (1.7)<br />

0 Id q<br />

In order to f<strong>in</strong>d a basis of left-<strong>in</strong>variant <strong>forms</strong> we compute A −1 dA with A ∈ U(p, q). If we<br />

denote A = (a 0 , . . . , a m ), then<br />

⎛<br />

⎞<br />

(da 0 , −a 0 ) . . . (da m , −a 0 )<br />

A −1 dA = εA T .<br />

.<br />

εdA =<br />

⎜ (da 0 , −a p−1 ) . . . (da m , −a p−1 )<br />

= (ϕ ij ) ij . (1.8)<br />

⎟<br />

⎝ (da 0 , a p ) . . . (da m , a p ) ⎠<br />

(da 0 , a m ) . . . (da m , a m )<br />

Each entry of this matrix is a 1-form given by ϕ ij = ±(da i , a j ).<br />

Note that each a j is a m-tuple of <strong>complex</strong> numbers, so that the 1-<strong>forms</strong> ϕ ij are <strong>complex</strong>valued.<br />

In order to f<strong>in</strong>d a basis of left-<strong>in</strong>variant real-valued 1-<strong>forms</strong> from the entries of the former<br />

matrix we use the follow<strong>in</strong>g<br />

• (a j , a j ) = ±1 and differentiat<strong>in</strong>g<br />

0 = (a j , da j ) + (da j , a j ) = (da j , a j ) + (da j , a j ) = 2Re(da j , a j ).<br />

Thus, ϕ jj = −ϕ jj and each ϕ jj takes only imag<strong>in</strong>ary values.<br />

• (a j , a k ) = 0 if k ≠ j and differentiat<strong>in</strong>g<br />

0 = (da j , a k ) + (a j , da k ).<br />

Thus, {<br />

ϕjk = ϕ kj if j ∈ {0, . . . , p − 1} or k ∈ {0, . . . , p − 1}<br />

ϕ jk = −ϕ kj<br />

otherwise.


1.2 Projective model 13<br />

On the other hand, it follows directly from the def<strong>in</strong>ition of U(p, q) that its dimension is<br />

(p + q) 2 . Then, {Re(ϕ jk ), Im(ϕ jk ), Re(ϕ jj )}, 0 ≤ j ≤ k ≤ m, j ≠ k constitutes a basis of<br />

left-<strong>in</strong>variant 1-<strong>forms</strong> s<strong>in</strong>ce they generate all the <strong>space</strong> and there are (p + q) 2 <strong>forms</strong>.<br />

For ɛ = 0, if we denote the elements A ∈ C n ⋊ U(n) as the matrices of (1.8), we def<strong>in</strong>e the<br />

<strong>forms</strong><br />

ϕ ij = (da i , a j )<br />

with ( , ) the standard product <strong>in</strong> C n+1 .<br />

Def<strong>in</strong>ition 1.2.8. A group G is said to be unimodular if there exists a volume element of G<br />

left and right-<strong>in</strong>variant.<br />

Lemma 1.2.9. U(p + q) is a unimodular group.<br />

Proof. From Lemma 1.2.7 we have a basis of left-<strong>in</strong>variant <strong>forms</strong> of U(p, q). We prove that<br />

each of these <strong>forms</strong> is also right-<strong>in</strong>variant, i.e. it satisfies<br />

R ∗ Bϕ ij (A; v) = ϕ ij (A; v)<br />

∀A, B ∈ U(p, q), v ∈ T A U(p, q).<br />

We use the expression ϕ ij = ±(da i , a j ) and we denote by a k (A) the map tak<strong>in</strong>g the k-th<br />

column of a matrix A. Then, if A, B ∈ U(p, q) and v ∈ T A U(p, q)<br />

RBϕ ∗ ij (A; v) = ϕ ij (R B (A); d(R B )(v)) = ±(da i (d(R B )(v)), a j (R B (A))) = ±(da i (vB), a j (AB))<br />

( m∑ ∑p−1<br />

m<br />

= ± − vi k b r k al j br l +<br />

∑ m∑<br />

)<br />

vi k b r k al j br l<br />

(<br />

= ± −<br />

(<br />

= ±<br />

k,l=0 r=0<br />

p−1<br />

∑<br />

k,l=0<br />

δ kl v k i a l j +<br />

m ∑<br />

k,l=p<br />

k,l=0 r=p<br />

δklv k i a l j<br />

∑p−1<br />

m<br />

− vi k a k j + ∑<br />

)<br />

vi k a k j<br />

= ±(da i (v), a j (A)) = ϕ ij (A; v).<br />

k=0<br />

k=p<br />

Then, U(p, q) is a unimodular group s<strong>in</strong>ce the volume element obta<strong>in</strong>ed from the product of the<br />

<strong>forms</strong> ϕ ij is left-<strong>in</strong>variant and right-<strong>in</strong>variant (it is a product of <strong>forms</strong> with this property).<br />

)<br />

1.2.6 Structure of homogeneous <strong>space</strong><br />

Def<strong>in</strong>ition 1.2.10. Let (M, g) be a Riemannian manifold. If given any two po<strong>in</strong>ts x, y ∈ M<br />

there exists an isometry σ of M such that σ(x) = y, then M is a homogeneous <strong>space</strong>. That is,<br />

a Riemannian manifold is homogeneous if it is a homogeneous <strong>space</strong> of its isometry group.<br />

By Proposition 1.2.6 we have that <strong>complex</strong> <strong>space</strong> <strong>forms</strong> are homogeneous <strong>space</strong>s. It will<br />

be <strong>in</strong>terest<strong>in</strong>g to represent them as a quotient of Lie groups.<br />

Proposition 1.2.11 ([War71] Theorem 3.62 page 123). Let η : G × M → M be a transitive<br />

action of the Lie group G over the manifold M. Let m 0 ∈ M and H the isotropy group of m 0 .<br />

Then, the map<br />

β : G/H −→ M<br />

gH ↦→ η(g, m 0 )<br />

is a diffeomorphism.


14 Spaces of constant holomorphic curvature<br />

The Lie group P U ɛ (n) acts transitively over CK n (ɛ) and the isotropy group of a po<strong>in</strong>t <strong>in</strong><br />

CK n (ɛ), for each ɛ, is isomorphic to P (U(1) × U(n)), a closed Lie subgroup of P U ɛ (n). Thus,<br />

we can represent CK n (ɛ) as a quotient of Lie groups<br />

CK n (ɛ) ∼ = P U ɛ (n)/P (U(1) × U(n)) ∼ = U ɛ (n)/(U(1) × U(n)),<br />

where the first diffeomorphism is given by x ↦→ g · P (U(1) × U(n)) with g ∈ P U ɛ (n) such that,<br />

if x 0 ∈ CK n (ɛ) is fixed then g(x 0 ) = x.<br />

Def<strong>in</strong>ition 1.2.12. A Riemannian manifold is a 2-po<strong>in</strong>t homogeneous <strong>space</strong> if the isometry<br />

group of the manifold acts transitively <strong>in</strong> the unit tangent bundle.<br />

Thus, <strong>complex</strong> <strong>space</strong> <strong>forms</strong> are 2-po<strong>in</strong>t homogeneous <strong>space</strong>s. Real <strong>space</strong> <strong>forms</strong> are also 2-<br />

po<strong>in</strong>t homogeneous <strong>space</strong>s but, they are also 3-po<strong>in</strong>t homogeneous <strong>space</strong>s, that is, the isometry<br />

group acts transitively for triplets of a po<strong>in</strong>t and two orthonormal vectors <strong>in</strong> the tangent <strong>space</strong><br />

of the po<strong>in</strong>t. Complex <strong>space</strong> <strong>forms</strong> (ɛ ≠ 0) cannot be 3-po<strong>in</strong>t homogeneous <strong>space</strong>s s<strong>in</strong>ce the<br />

sectional curvature is preserved by isometries and the sectional curvature is not constant.<br />

1.3 Mov<strong>in</strong>g frames<br />

Def<strong>in</strong>ition 1.3.1. Let U ⊂ M be an open set of a differentiable manifold. An orthonormal<br />

mov<strong>in</strong>g frame of CK n (ɛ) def<strong>in</strong>ed at U is a map g 0 : U → CK n (ɛ) together with a collection of<br />

g i : U → T CK n (ɛ) (i ∈ {1, . . . , 2n}) such that 〈g i , g j 〉 ɛ = δ ij where 〈 , 〉 ɛ denotes the Hermitian<br />

product of CK n (ɛ) (def<strong>in</strong>ed at (1.4)) and π : T CK n (ɛ) → CK n (ɛ) is the canonical projection.<br />

Def<strong>in</strong>ition 1.3.2. Let V be a 2n-dimensional real vector <strong>space</strong> endowed with a <strong>complex</strong><br />

structure J. An orthonormal basis {v 1 , v 2 , . . . , v 2n } is said to be a J-basis if v 2i = Jv 2i−1 for<br />

every i ∈ {1, . . . , n}.<br />

We denote by {e 1 , e 1 = Je 1 , . . . , e n , e n = Je n } the J-bases of V , and by {ω 1 , ω 1 , . . . , ω n , ω n }<br />

the dual basis of a J-basis.<br />

Remark 1.3.3. A J-basis is a special type of an orthonormal basis <strong>in</strong> a real vector <strong>space</strong> with<br />

an almost <strong>complex</strong> structure J.<br />

Def<strong>in</strong>ition 1.3.4. An orthonormal mov<strong>in</strong>g frame of CK n (ɛ) such that vectors {g 1 (p), . . . , g 2n (p)}<br />

constitute a J-basis for all x ∈ U, is called a J-mov<strong>in</strong>g frame.<br />

J-mov<strong>in</strong>g frames <strong>in</strong> CK n (ɛ) play an important role s<strong>in</strong>ce they are <strong>in</strong> correspondence with<br />

the elements of the isometry group of CK n (ɛ).<br />

Consider F(CK n (ɛ)) the bundle of J-mov<strong>in</strong>g frames of CK n (ɛ), constituted by J-mov<strong>in</strong>g<br />

frames (g 0 ; g 1 , Jg 1 , . . . , g n , Jg n ) with g 0 ∈ CK n (ɛ) and {g 1 , Jg 1 , . . . , g n , Jg n } a J-basis of<br />

T g0 CK n (ɛ).<br />

Proposition 1.3.5. The bundle of J-mov<strong>in</strong>g frames F(CK n (ɛ)) is identified with the isometry<br />

group of CK n (ɛ).<br />

Proof. We study the case ɛ ≠ 0. Let A ∈ U ɛ (n). By def<strong>in</strong>ition (1.5) of U ɛ (n) we have that A is<br />

an (n + 1) × (n + 1) matrix with <strong>complex</strong> entries and such that its columns {a 0 , . . . , a n } satisfy<br />

⎧<br />

⎨<br />

⎩<br />

(a 0 , a 0 ) = sign(ɛ)1,<br />

(a 0 , a i ) = 0, i ∈ {1, . . . , n},<br />

(a i , a j ) = δ ij i, j ∈ {1, . . . , n},<br />

where ( , ) denotes the Hermitian product <strong>in</strong> C n+1 def<strong>in</strong>ed at (1.2).<br />

(1.9)


1.3 Mov<strong>in</strong>g frames 15<br />

From the first property <strong>in</strong> the former list, we can take a 0 as a representative of g 0 = π(a 0 ) ∈<br />

CK n (ɛ).<br />

From the second property of (1.9) we have that a i satisfies the condition of be<strong>in</strong>g a vector<br />

<strong>in</strong> T a0 H (cf. Section 1.2.2). Moreover, Re(Ja 0 , a i ) = Im(a 0 , a i ) = 0 and Re(Ja 0 , Ja i ) =<br />

Re(a 0 , a i ) = 0. Let us consider g i := dπ(a i ) and Jg i := dπ(Ja i ) where π denotes the projection<br />

def<strong>in</strong>ed at (1.3).<br />

From the third condition <strong>in</strong> (1.9), vectors {g 1 , Jg 1 , . . . , g n , Jg n } constitute a J-basis of the<br />

tangent <strong>space</strong> at g 0 .<br />

Reciprocally, given a J-mov<strong>in</strong>g frame {g 0 ; g 1 , Jg 1 , . . . , g n , Jg n } def<strong>in</strong>ed on an open set, we<br />

can def<strong>in</strong>e a matrix of U ɛ (n) (with the entries depend<strong>in</strong>g cont<strong>in</strong>uously on a parameter) just<br />

tak<strong>in</strong>g as the first column the representative a 0 of g 0 with norm sign(ɛ)1. For the other columns<br />

we consider the horizontal lift of g j at a 0 . As {g 1 , Jg 1 , . . . , g n , Jg n } is, <strong>in</strong> each po<strong>in</strong>t g, a J-basis<br />

and we choose the horizontal lift, the columns of the constructed matrix verify the conditions<br />

<strong>in</strong> (1.9) and are <strong>in</strong> U ɛ (n).<br />

Def<strong>in</strong>ition 1.3.6. The unit tangent bundle of CK n (ɛ), denoted by S(CK n (ɛ)), is def<strong>in</strong>ed as<br />

S(CK n (ɛ)) =<br />

⋃<br />

p∈CK n (ɛ)<br />

T ′ pCK n (ɛ)<br />

where T ′ pCK n (ɛ) denotes the sphere of unit vectors <strong>in</strong> the tangent vector <strong>space</strong> of CK n (ɛ) at<br />

p.<br />

In Lemma 1.2.7 we def<strong>in</strong>ed the <strong>in</strong>variant <strong>forms</strong> {ϕ ij } of U ɛ (n) as<br />

ϕ ij (A; ·) = (da i (·), a j )<br />

where A = (a 0 , . . . , a n ) ∈ U ɛ (n). As <strong>forms</strong> {ϕ ij } takes <strong>complex</strong> values we consider<br />

ϕ jk = α jk + iβ jk (1.10)<br />

Us<strong>in</strong>g the identification between U ɛ (n) and F(CK n (ɛ)), we can consider <strong>forms</strong> {ϕ ij } as<br />

<strong>forms</strong> of F(CK n (ɛ)).<br />

On the other hand, consider the canonical projections<br />

and local sections<br />

π 1<br />

F(CK n (ɛ)) −→ S(CK n (ɛ)) −→ CK n (ɛ)<br />

(g; g 1 , . . . , Jg n ) ↦→ (g, g 1 ) ↦→ g<br />

CK n (ɛ) ⊃ U s 2<br />

−→ S(CK n (ɛ)) ⊃ V s 1<br />

−→ F(CK n (ɛ)).<br />

Us<strong>in</strong>g the <strong>forms</strong> ϕ ij def<strong>in</strong>ed <strong>in</strong> F(CK n (ɛ)) and the previous local sections, we def<strong>in</strong>e the<br />

follow<strong>in</strong>g local <strong>in</strong>variant <strong>forms</strong> <strong>in</strong> S(CK n (ɛ))<br />

and the local <strong>in</strong>variant <strong>forms</strong> <strong>in</strong> CK n (ɛ)<br />

s ∗ 1(ϕ ij ), s ∗ 1(α ij ) and s ∗ 1(β ij )<br />

s ∗ 2s ∗ 1(ϕ ij ), s ∗ 2s ∗ 1(α ij ) and s ∗ 2s ∗ 1(β ij ).<br />

Lemma 1.3.7. Forms s ∗ 1 α 01, s ∗ 1 β 01 and s ∗ 1 β 11 are global <strong>forms</strong> <strong>in</strong> S(CK n (ɛ)).<br />

π 2


16 Spaces of constant holomorphic curvature<br />

Proof. If V ∈ T (p,v) (S(CK n (ɛ))) then<br />

s ∗ 1α 01 (V ) (p,v) = Re((dπ 2 (V ), v)) = 〈dπ 2 (V ), v〉 ɛ<br />

s ∗ 1β 01 (V ) (p,v) = Im((dπ 2 (V ), v)),<br />

s ∗ 1β 11 (V ) (p,v) = Im((∇V, v)),<br />

where ∇ denotes the Levi-Civita connection def<strong>in</strong>ed by ∇ : T (SCK n (ɛ)) → T CK n (ɛ), from<br />

the Levi-Civita connection of CK n (ɛ). Inded, a vector V ∈ T (SCK n (ɛ)) is a tangent vector<br />

to a curve of unit vectors, and <strong>in</strong> each of them we can apply the Levi-Civita connection of<br />

CK n (ɛ).<br />

We denote by α, β, γ the <strong>forms</strong> s ∗ 1 α 01, s ∗ 1 β 01, s ∗ 1 β 11, respectively.<br />

Remark 1.3.8. The 1-form α co<strong>in</strong>cides with the standard contact form of the unit tangent<br />

bundle S(CK n (ɛ)).<br />

On the other hand, <strong>forms</strong> s ∗ 2 s∗ 1 (ϕ ij) co<strong>in</strong>cide with <strong>forms</strong> φ ij of CK n (ɛ) we def<strong>in</strong>e <strong>in</strong> the<br />

follow<strong>in</strong>g.<br />

Let {g; g 1 , Jg 1 , . . . , g n , Jg n } be a J-mov<strong>in</strong>g frame on CK n (ɛ). As <strong>in</strong> the tangent <strong>space</strong> of<br />

each po<strong>in</strong>t g, {g 1 , Jg 1 , . . . , g n , Jg n } def<strong>in</strong>es a J-basis, we can consider the vectors {g 1 , . . . , g n }<br />

as <strong>complex</strong> vectors. Then, the follow<strong>in</strong>g differential <strong>forms</strong> are well-def<strong>in</strong>ed<br />

φ j (·) = (dg(·), g j ) ɛ and φ jk (·) = (∇g j (·), g k ) ɛ (1.11)<br />

where j, k ∈ {1, ..., n}, and ∇ denotes the Levi-Civita of CK n (ɛ) (i.e. we consider g j as a real<br />

vector, we apply the Levi-Civita connection and we consider the result aga<strong>in</strong> as a <strong>complex</strong><br />

vector). Note that the differential <strong>forms</strong> φ j and φ jk are <strong>complex</strong> valued. We denote<br />

φ j = α j + iβ j (1.12)<br />

φ jk = α jk + iβ jk .<br />

At Chapter 3, we work with orthonormal mov<strong>in</strong>g frames not necessarily J-mov<strong>in</strong>g frames.<br />

Analogously, if {g; g 1 , g 2 , . . . , g 2n−1 , g 2n } is a mov<strong>in</strong>g frame on CK n (ɛ), we def<strong>in</strong>e the dual and<br />

connection <strong>forms</strong> for this mov<strong>in</strong>g frame. We denote<br />

ω j (·) = 〈dg(·), g j 〉 ɛ and ω jk (·) = 〈∇g j (·), g k 〉 ɛ (1.13)<br />

with j, k ∈ {1, ..., 2n}, 〈 , 〉 ɛ the Hermitian product def<strong>in</strong>ed on CK n (ɛ) (see (1.4)), and ∇ the<br />

Levi-Civita connection of CK n (ɛ).<br />

Note that the differential <strong>forms</strong> {α j , β j } are a particular case of <strong>forms</strong> {ω j }: they are<br />

obta<strong>in</strong>ed if we consider a J-mov<strong>in</strong>g frame.<br />

Notation 1.3.9. Along this work we use <strong>in</strong>variant <strong>forms</strong> def<strong>in</strong>ed at CK n (ɛ), S(CK n (ɛ)) or<br />

F(CK n (ɛ)), but we denote all of them by ϕ ij , α ij , β ij , without the pull-back of the sections,<br />

if it is clear by the context.<br />

Def<strong>in</strong>ition 1.3.10. Given a doma<strong>in</strong> Ω ⊂ CK n (ɛ) we def<strong>in</strong>e the unit normal bundle of ∂Ω by<br />

N(Ω)={(p, v) : p ∈ ∂Ω, v such that 〈v, w〉 ɛ ≥ 0 ∀ w tangent to a curve at Ω by p and ||v|| ɛ =1}.<br />

Remark 1.3.11. The ma<strong>in</strong> results of this work, given at Chapters 3 and 4, have as a hypothesis<br />

that the doma<strong>in</strong> Ω ⊂ CK n (ɛ) which we take is compact with C 2 boundary. We denote by<br />

regular doma<strong>in</strong> a doma<strong>in</strong> satisfy<strong>in</strong>g these hypothesis. We suppose that doma<strong>in</strong>s are regular <strong>in</strong><br />

order to simplify the arguments and to use techniques of differential <strong>geometry</strong> (for <strong>in</strong>stance,<br />

to have a well-def<strong>in</strong>ed second fundamental form <strong>in</strong> the whole boundary of the doma<strong>in</strong>). These<br />

hypothesis can be relaxed s<strong>in</strong>ce most of the used results, ma<strong>in</strong>ly <strong>in</strong> valuations (see Chapter 2),<br />

are known for a more general class of doma<strong>in</strong>s (cf. [Ale07a]).


1.4 Submanifolds 17<br />

Lemma 1.3.12. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>. Forms α and dα vanishes at N(Ω) ⊂<br />

S(CK n (ɛ)).<br />

Proof. Let V ∈ T (p,v) Ω. Then, α(V ) (p,N) = 〈dπ 2 (V ), N〉 ɛ = 0 s<strong>in</strong>ce dπ 2 (V ) is a tangent vector<br />

tangent at ∂Ω at po<strong>in</strong>t p.<br />

In order to prove that the 2-form dα vanishes at the unit normal bundle, we consider the<br />

<strong>in</strong>clusion of the unit normal bundle to the unit tangent bundle i : N(Ω) → S(Ω). Us<strong>in</strong>g that<br />

the differential map commutes with the <strong>in</strong>clusion we have the result<br />

dα| N(Ω) = (i ∗ ◦ d)(α) = (d ◦ i ∗ )(α) = d(i ∗ α) = d(0) = 0.<br />

1.4 Submanifolds<br />

1.4.1 Totally geodesic submanifolds<br />

Def<strong>in</strong>ition 1.4.1. Let M be a Riemannian manifold. A submanifold N ⊂ M is totally geodesic<br />

if every geodesic <strong>in</strong> the submanifold N is also a geodesic <strong>in</strong> M.<br />

As C n is metrically equivalent to R 2n , totally geodesic submanifolds <strong>in</strong> C n co<strong>in</strong>cide with<br />

the ones <strong>in</strong> R 2n . For the other <strong>complex</strong> <strong>space</strong> <strong>forms</strong>, the totally geodesic submanifolds are<br />

classified.<br />

Def<strong>in</strong>ition 1.4.2. Let V be a real vector <strong>space</strong> of dimension 2n endowed with an almost <strong>complex</strong><br />

structure J compatible with a scalar product 〈 , 〉. It is said that vectors {e 1 , . . . , e m } expand<br />

a <strong>complex</strong> sub<strong>space</strong> if the <strong>space</strong> generated by these vectors is J-<strong>in</strong>variant, i.e. J(span{e 1 , . . . , e m }) =<br />

span{e 1 , . . . , e m }.<br />

It is said that a submanifold of a <strong>complex</strong> manifold is <strong>complex</strong> if at each po<strong>in</strong>t, the tangent<br />

<strong>space</strong> of the submanifold is <strong>complex</strong> sub<strong>space</strong> of the tangent <strong>space</strong> of the manifold.<br />

Def<strong>in</strong>ition 1.4.3. Let V be a real vector <strong>space</strong> of dimension 2n endowed with an almost<br />

<strong>complex</strong> structure J compatible with a scalar product 〈 , 〉. It is said that vectors {e 1 , . . . , e m }<br />

expand a totally real sub<strong>space</strong> if<br />

〈e i , Je j 〉 = 0, ∀i, j ∈ {1, . . . , m}.<br />

It is said that a submanifold of a <strong>complex</strong> manifold is totally real if at each po<strong>in</strong>t, the tangent<br />

<strong>space</strong> of the submanifold is a totally real sub<strong>space</strong> of the tangent <strong>space</strong> of the manifold.<br />

Theorem 1.4.4 ([Gol99] pages 75 and 80). Let z ∈ CK n (ɛ).<br />

1. If L ⊂ T z CK n (ɛ) is a <strong>complex</strong> vector sub<strong>space</strong> with <strong>complex</strong> dimension r, then there<br />

exists a unique complete <strong>complex</strong> totally geodesic submanifold through z and tangent to<br />

L at z.<br />

2. If L ⊂ T z CK n (ɛ) is a totally real vector <strong>space</strong> of real dimension k, then there exists a<br />

unique complete totally geodesic totally real submanifold through z and tangent to L at<br />

z.<br />

Def<strong>in</strong>ition 1.4.5. The <strong>complex</strong> submanifold def<strong>in</strong>ed at 1. <strong>in</strong> the previous theorem is called<br />

<strong>complex</strong> r-plane, and denoted by L r .<br />

The totally real submanifold def<strong>in</strong>ed at 2. <strong>in</strong> the previous theorem is called totally real<br />

k-plane, and denoted by L R k .


18 Spaces of constant holomorphic curvature<br />

In the projective model, <strong>complex</strong> r-planes are obta<strong>in</strong>ed from the projection of a sub<strong>space</strong><br />

F ⊂ C n+1 <strong>in</strong>tersection CK n (ɛ). The sub<strong>space</strong> F is the (r + 1)-dimensional <strong>complex</strong> vector<br />

sub<strong>space</strong> spanned by a representative z ′ of z = π(z ′ ) and by the horizontal lift of vectors <strong>in</strong><br />

L ⊂ T z CK n (ɛ) (cf. [Gol99, Section 3.1.4]).<br />

Analogously, totally real k-planes are obta<strong>in</strong>ed from the projection of F R ⊂ C n+1 <strong>in</strong>tersection<br />

CK n (ɛ), where F R is the (k + 1)-dimensional real vector sub<strong>space</strong> spanned by a<br />

representative z ′ of z = π(z ′ ) and by the horizontal lift of vectors <strong>in</strong> L ⊂ T z CK n (ɛ).<br />

Theorem 1.4.6 ([Gol99] p. 82). The unique complete totally geodesic submanifolds <strong>in</strong> CK n (ɛ)<br />

are the <strong>complex</strong> r-planes, r ∈ {1, . . . , n − 1} and the totally real k-planes, k ∈ {1, . . . , n}.<br />

Corollary 1.4.7. In CK n (ɛ), ɛ ≠ 0, there are not totally geodesic (real) hypersurfaces.<br />

That is, it does not exist the equivalent hypersurface to a hyperplane <strong>in</strong> a real <strong>space</strong> form.<br />

The more reasonable substitutes of hyperplanes are the so-called bisectors, which we study on<br />

page 82.<br />

Theorem 1.4.6 will be important along this work s<strong>in</strong>ce it will be <strong>in</strong>terest<strong>in</strong>g to know which<br />

totally geodesic submanifolds can be taken <strong>in</strong> a <strong>complex</strong> <strong>space</strong> form as a substitutes of (totally<br />

geodesic) planes <strong>in</strong> real <strong>space</strong> <strong>forms</strong>.<br />

1.4.2 Geodesic balls<br />

A geodesic ball <strong>in</strong> a Riemannian manifold is the set of po<strong>in</strong>ts equidistant from a fixed po<strong>in</strong>t<br />

called center.<br />

In real <strong>space</strong> <strong>forms</strong>, geodesic balls are totally umbilical real hypersurfaces, i.e. the second<br />

fundamental form is, at every po<strong>in</strong>t, a multiple of the identity and the same multiple for every<br />

po<strong>in</strong>t.<br />

This fact does not hold <strong>in</strong> <strong>complex</strong> projective and hyperbolic <strong>space</strong>s. Moreover, <strong>in</strong> these<br />

<strong>space</strong>s, there are no totally umbilical real hypersurface.<br />

Proposition 1.4.8 ([Mon85]). The pr<strong>in</strong>cipal curvatures of a sphere of radius r <strong>in</strong> CK n (ɛ),<br />

ɛ ≠ 0 are<br />

i) 2 cot ɛ (2r) with multiplicity 1 and pr<strong>in</strong>cipal direction −JN (where N denotes the <strong>in</strong>ward<br />

normal vector to the sphere),<br />

ii) cot ɛ (r) with multiplicity 2n − 2.<br />

Recall that cos ɛ , s<strong>in</strong> ɛ denote the generalized trigonometric functions def<strong>in</strong>ed at Notation<br />

1.2.2.<br />

Along this work, we use the value of the mean curvature <strong>in</strong>tegrals for a geodesic ball of<br />

radius R <strong>in</strong> CK n (ɛ) (cf. Def<strong>in</strong>ition 2.1.4).<br />

From the previous proposition we have that the symmetric elementary functions are<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

σ 0 = 1<br />

) −1<br />

( (2n−2<br />

σ i = ( 2n−1<br />

i<br />

σ 2n−1 = 2 cot 2n−2<br />

ɛ<br />

i<br />

)<br />

cot<br />

i<br />

ɛ (R) + ( 2n−2<br />

i−1<br />

(R) cot ɛ (2R).<br />

) )<br />

2 cot<br />

i−1(R) cot ɛ (2R)<br />

By a straightforward computation, we obta<strong>in</strong> that the expression of the mean curvature <strong>in</strong>tegrals<br />

is<br />

and<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

M 0 = vol(∂B R ) =<br />

2π<br />

M i =<br />

n<br />

(2n−1)(n−1)!<br />

M 2n−1 =<br />

2πn<br />

(n−1)! (cos2n ɛ<br />

2πn<br />

(n−1)! s<strong>in</strong>2n−1 ɛ<br />

((2n + i − 1) cosi+1 ɛ<br />

(R) + cos 2n−2<br />

ɛ<br />

(R) cos ɛ (R)<br />

(R) s<strong>in</strong> 2n−i−1<br />

ɛ<br />

(R) s<strong>in</strong> 2 ɛ(R))<br />

vol(B R ) = πn<br />

n! (s<strong>in</strong> ɛ(R)) 2n .<br />

ɛ<br />

(R) − i cos i−1 (R) s<strong>in</strong> 2n−i−1 (R))<br />

ɛ<br />

ɛ


1.5 Space of <strong>complex</strong> r-planes 19<br />

1.5 Space of <strong>complex</strong> r-planes<br />

We denote the <strong>space</strong> of all <strong>complex</strong> r-planes <strong>in</strong> CK n (ɛ) by L C r and the <strong>space</strong> of all totally real<br />

r-planes <strong>in</strong> CK n (ɛ) by L R r (cf. Def<strong>in</strong>ition 1.4.5).<br />

We shall use that L C r is a homogeneous <strong>space</strong> with respect to the isometry group of CK n (ɛ).<br />

In order to prove this fact, we need the follow<strong>in</strong>g result.<br />

Lemma 1.5.1. The group of isometries of CK n (ɛ) acts transitively on the J-bases.<br />

Proof. We study the case ɛ ≠ 0. Fix the canonical J-basis {e 1 , Je 1 , . . . , e n , Je n } at e 0 ∈<br />

CK n (ɛ). It is enough to prove that given another J-basis {g 1 , Jg 1 , . . . , g n , Jg n } of T g0 CK n (ɛ),<br />

there exists an isometry ρ which takes this J-basis to the fixed one.<br />

Take as isometry ρ ∈ U ɛ (n) the matrix with columns (˜g 0 , ˜g 1 , . . . , ˜g n ) where ˜g 0 is a representative<br />

of g 0 with norm ɛ and ˜g i is the horizontal lift of g i at ˜g 0 . In the same way as <strong>in</strong> the<br />

proof of Proposition 1.3.5 we have that ρ is a matrix <strong>in</strong> U ɛ (n). Moreover, it carries the fixed<br />

J-basis to the given one.<br />

From the previous lemma we get<br />

Lemma 1.5.2. The <strong>space</strong> of <strong>complex</strong> r-planes is a homogeneous <strong>space</strong> with respect to the<br />

isometry group of CK n (ɛ).<br />

Proof. We def<strong>in</strong>e a J-basis {e 1 , Je 1 , . . . , e r , Je r } of the tangent <strong>space</strong> <strong>in</strong> any po<strong>in</strong>t of a <strong>complex</strong><br />

r-plane. Complet<strong>in</strong>g this J-basis to a J-basis of the whole <strong>space</strong> CK n (ɛ), and apply<strong>in</strong>g the<br />

previous lemma we get the result.<br />

In order to study <strong>in</strong>tegral <strong>geometry</strong> <strong>in</strong> CK n (ɛ), it is necessary that the <strong>space</strong> L C r admits an<br />

<strong>in</strong>variant density under the isometries of CK n (ɛ). In general, the absolute value of a form of<br />

maximum degree is called a density. By the follow<strong>in</strong>g lemma, it is enough to prove that L C r is<br />

the quotient of unimodular groups.<br />

Lemma 1.5.3 ([San04]). If G, H are unimodular groups, then G/H admits an <strong>in</strong>variant<br />

density.<br />

The isotropy group of a <strong>complex</strong> r-plane is isomorphic to<br />

U ɛ (r) × U(n − r) =<br />

{<br />

( A 0<br />

X ∈ M (n+1)×(n+1) (C) : X =<br />

0 B<br />

)<br />

}<br />

, A ∈ U ɛ (r), B ∈ U(n − r)<br />

(1.14)<br />

s<strong>in</strong>ce these matrices (and only these) leave <strong>in</strong>variant a <strong>complex</strong> r-plane and its orthogonal.<br />

Then,<br />

L C r ∼ = U ɛ (n)/(U ɛ (r) × U(n − r)).<br />

The group U ɛ (n), by Lemma 1.2.9, is unimodular, and U ɛ (r)×U(n−r) is also a unimodular<br />

group. Thus, there exists an <strong>in</strong>variant density <strong>in</strong> the quotient <strong>space</strong>, that is, <strong>in</strong> the <strong>space</strong> of<br />

<strong>complex</strong> r-planes.<br />

The follow<strong>in</strong>g result give a method to obta<strong>in</strong> explicitly the density <strong>in</strong> the quotient <strong>space</strong>.<br />

Theorem 1.5.4 ([San04] page 147). Let G be a Lie group with dimension n and H a closed<br />

subgroup of G with dimension n − m. Then, G/H is a homogeneous <strong>space</strong> with dimension m.<br />

Let ˜ω be the m-form obta<strong>in</strong>ed from the product of all <strong>in</strong>variant 1-<strong>forms</strong> of G such that they<br />

vanish on H. Then, there exists an <strong>in</strong>variant density ω <strong>in</strong> G/H if and only if d˜ω = 0. In<br />

this case, ˜ω is the pull-back of ω for the canonical projection of G at G/H (up to constants<br />

factors).


20 Spaces of constant holomorphic curvature<br />

Proposition 1.5.5. Let π : U ɛ (n) −→ U ɛ (n)/(U ɛ (r) × U(n − r)) ∼ = L C r . If dL r is an <strong>in</strong>variant<br />

density of L C r then<br />

∧<br />

π ∗ dL r = ϕ ij ∧ ϕ ij<br />

i=0,...,r<br />

j=r+1,...,n<br />

or, equivalently,<br />

π ∗ dL r =<br />

∧<br />

i=0,...,r<br />

j=r+1,...,n<br />

α ij ∧ β ij ,<br />

where {ϕ ij , α ij , β ij } are the <strong>forms</strong> def<strong>in</strong>ed at Lemma 1.2.7 and at (1.11).<br />

Proof. The result is known for C n , see [San04].<br />

For the other <strong>complex</strong> <strong>space</strong> <strong>forms</strong>, we note that the (real) dimension of the <strong>space</strong> of<br />

<strong>complex</strong> r-plans co<strong>in</strong>cides with the degree of π ∗ dL r (and it is equal to 2(r + 1)(n − r)). Thus,<br />

it suffices to prove that each form ϕ ij with i ∈ {0, . . . , r}, j ∈ {r + 1, . . . , n} vanishes over<br />

U ɛ (n) × U(n − r). But, from the form of matrices <strong>in</strong> U ɛ (n) × U(n − r) given at (1.14), the<br />

result follows immediately.<br />

Let us give an example of mov<strong>in</strong>g frames <strong>in</strong> the <strong>space</strong> of <strong>complex</strong> r-planes us<strong>in</strong>g Def<strong>in</strong>ition<br />

1.3.1, which will be used <strong>in</strong> the next section.<br />

Let us take as the open set U ⊂ M (see Def<strong>in</strong>ition 1.3.1) an open set <strong>in</strong> L C r . An orthonormal<br />

frame is given by<br />

g : U ⊂ L C r −→ CK n (ɛ)<br />

L r ↦→ p ∈ L r<br />

and<br />

g i : U ⊂ L C r −→ T CK n (ɛ)<br />

L r ↦→ v i ∈ T g(Lr)L r<br />

, (1.15)<br />

i ∈ {1, . . . , 2r}, such that 〈v i , v j 〉 ɛ = δ ij . It will be <strong>in</strong>terest<strong>in</strong>g to consider that Jg 2k−1 (L r ) =<br />

g 2k (L r ), k ∈ {1, . . . , r}, that is, vectors {g 1 (L r ), . . . , g 2r (L r )} constitute a J-basis at g(L r ). By<br />

abuse of notation, we denote g(L r ) by g and g i (L r ) by g i .<br />

Remark 1.5.6. From the correspondence between U ɛ (n) and the bundle of J-mov<strong>in</strong>g frames<br />

F(CK n (ɛ)), and between U ɛ (n)/(U ɛ (r)×U(n−r)) and L C r we have that {p; g 1 (L r ), . . . , g 2r (L r )}<br />

are sections of U ɛ (n) → U ɛ (n)/(U ɛ (r) × U(n − r)).<br />

1.5.1 Expression for the <strong>in</strong>variant density <strong>in</strong> terms of a parametrization<br />

Sometimes it is <strong>in</strong>terest<strong>in</strong>g to have a more geometric expression for the <strong>in</strong>variant density of<br />

<strong>complex</strong> r-planes. For example, Santaló proved<br />

Proposition 1.5.7 ([San04]). The <strong>in</strong>variant density of the <strong>space</strong> of totally geodesic planes <strong>in</strong><br />

a <strong>space</strong> form of constant sectional curvature k is<br />

dL r = cos r k (ρ)dx n−r ∧ dL (n−r)[O] (1.16)<br />

where dx n−r is the volume element of the (n − r)-plane orthogonal to L r conta<strong>in</strong><strong>in</strong>g the orig<strong>in</strong><br />

O and dL (n−r)[O] is the volume element of the Grassmannian of (n − r)-planes conta<strong>in</strong><strong>in</strong>g the<br />

orig<strong>in</strong>.<br />

Now, we give a similar expression <strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong>, for the density of <strong>complex</strong><br />

r-planes <strong>in</strong> CK n (ɛ).


1.5 Space of <strong>complex</strong> r-planes 21<br />

Proposition 1.5.8. The <strong>in</strong>variant density of the <strong>space</strong> of <strong>complex</strong> r-planes <strong>in</strong> a <strong>space</strong> form<br />

of constant holomorphic curvature 4ɛ is<br />

dL r = cos 2r<br />

ɛ (ρ)dx n−r ∧ dL (n−r)[O]<br />

where dx n−r is the volume element of the <strong>complex</strong> (n−r)-plane orthogonal to L r conta<strong>in</strong><strong>in</strong>g the<br />

orig<strong>in</strong> O and dL (n−r)[O] is the volume element of the Grassmannian of <strong>complex</strong> (n − r)-planes<br />

conta<strong>in</strong><strong>in</strong>g the orig<strong>in</strong>.<br />

Proof. In order to obta<strong>in</strong> the expression of the <strong>in</strong>variant density of L C r <strong>in</strong> terms of dL (n−r)[O]<br />

and dx n−r , we follow the same idea as <strong>in</strong> [San04] (where it is proved for the Euclidean and the<br />

real hyperbolic <strong>space</strong>). That is, we fix an adapted J-mov<strong>in</strong>g frame to the <strong>complex</strong> r-plane,<br />

def<strong>in</strong>ed <strong>in</strong> a neighborhood of the po<strong>in</strong>t at m<strong>in</strong>imum distance from the orig<strong>in</strong> O, then, by<br />

parallel translation, we translate this mov<strong>in</strong>g frame to the orig<strong>in</strong> O. F<strong>in</strong>ally, we relate both<br />

mov<strong>in</strong>g frames from the pull-back of a section.<br />

Denote by G = U ɛ (n) and by H = U ɛ (r)×U(n−r) the isotropy group of a <strong>complex</strong> r-plane.<br />

The projection π : G −→ G/H gives, with respect to a J-mov<strong>in</strong>g frame adapted to the<br />

<strong>complex</strong> r-plane and the <strong>forms</strong> def<strong>in</strong>ed at Lemma 1.2.7,<br />

π ∗ dL r =<br />

n∧<br />

j=r+1<br />

α j0 ∧ β j0<br />

∧<br />

i=1,...,r<br />

j=r+1,...,n<br />

α ji ∧ β ji .<br />

Let O ∈ CK n (ɛ) be the fixed orig<strong>in</strong> and let L r ∈ L C r . We denote by p(L r ) the po<strong>in</strong>t <strong>in</strong> L r<br />

at a m<strong>in</strong>imum distance from O.<br />

Let i be a local section of π. Then, π ◦ i = id and i ∗ π ∗ dL r = dL r , so that, we can obta<strong>in</strong><br />

the expression of dL r . From the identifications expla<strong>in</strong>ed <strong>in</strong> Remark 1.5.6, we take as a section<br />

i the def<strong>in</strong>ed by {p(L r ); g 1 , Jg 1 , . . . , g n , Jg n }, a J-mov<strong>in</strong>g frame def<strong>in</strong>ed <strong>in</strong> a neighborhood V<br />

of L r , adapted to L r and such that g r+1 is the tangent vector to the geodesic jo<strong>in</strong><strong>in</strong>g p(L r )<br />

and O. Denote by {g 1 , Jg 1 , . . . , g n , Jg n } the dual basis of {g 1 , Jg 1 , . . . , g n , Jg n } at g 0 . From<br />

Proposition 1.3.5, we consider the matrix <strong>in</strong> G correspond<strong>in</strong>g to this J-mov<strong>in</strong>g frame. Denote<br />

the columns of G also by (g 0 , g 1 , . . . , g n ), so that (g i ◦ i) denotes the i-th column of the matrix.<br />

Then, us<strong>in</strong>g the same notation as <strong>in</strong> (1.12), we have<br />

n∧<br />

i ∗ ( α j0 ∧ β j0 ) =<br />

j=r+1<br />

=<br />

n∏<br />

n∏<br />

α j0 (di) ∧ β j0 (di) = 〈dg 0 (di), g j 〉〈dg 0 (di), Jg j 〉<br />

j=r+1<br />

n∏<br />

j=r+1<br />

〈d(g 0 ◦ i), g j 〉〈d(g 0 ◦ i), Jg j 〉 = (g 0 ◦ i) ∗ (<br />

j=r+1<br />

j=r+1<br />

n∧<br />

g j ∧ Jg j )<br />

but (g 0 ◦ i) = p(L r ) and the previous form co<strong>in</strong>cides with the volume element of p <strong>in</strong> the<br />

sub<strong>space</strong> generated by {g r+1 , Jg r+1 , . . . , g n , Jg n }, which is a <strong>complex</strong> (n − r)-plane. Thus,<br />

i ∗ (<br />

n∧<br />

j=r+1<br />

α j0 ∧ β j0 ) = dx n−r .<br />

Let G ′ ⊂ G be the subgroup of all J-mov<strong>in</strong>g frames of G such that g r+1 is the tangent<br />

vector to the geodesic conta<strong>in</strong><strong>in</strong>g O and g 0 , and let G 0 ⊂ G be the subgroup of all J-mov<strong>in</strong>g<br />

frames with base po<strong>in</strong>t O. Let ρ be the distance from L r to O. Denote by s ρ the parallel<br />

translation from O to p(L r ) along the geodesic. Consider the follow<strong>in</strong>g maps<br />

π 1 : G ′ −→ G 0<br />

(g 0 ; g 1 , . . . , g n ) ↦→ (O; s −1<br />

ρ (g 1 ), . . . , s −1<br />

ρ (g n )) ,


22 Spaces of constant holomorphic curvature<br />

π 2 : G 0 −→ L C r[O]<br />

G 0 ↦→ <strong>complex</strong> r-plane conta<strong>in</strong><strong>in</strong>g O generated by g 1 , . . . , g r<br />

,<br />

π 3 : L C r −→ L C r[0]<br />

L r ↦→ ((L r ) ⊥ O )⊥ O<br />

where (L r ) ⊥ O denotes the orthogonal <strong>space</strong> to L r by O. Then, the follow<strong>in</strong>g diagram is commutative<br />

G ′ π −−−−→<br />

1<br />

G0<br />

⏐ ⏐<br />

π↓<br />

↓ π 2<br />

(1.17)<br />

L C r<br />

We def<strong>in</strong>e curves x ij <strong>in</strong> G 0 ⊂ G as follows<br />

π 3<br />

−−−−→ L<br />

C<br />

r[0]<br />

.<br />

x ij (t) = (O; g 1 , . . . , g i (t), Jg i (t), . . . , g j (t), Jg j (t) . . . , g n , Jg n )<br />

where g i (t) = cos(t)g i + s<strong>in</strong>(t)g j and g j (t) = − s<strong>in</strong>(t)g i + cos(t)g j and curves<br />

x j i (t) = (O; g 1, . . . , g i (t), Jg i (t), . . . , g j (t), Jg j (t), . . . , g n , Jg n )<br />

where g i (t) = cos(t)g i + s<strong>in</strong>(t)Jg j and g j (t) = − s<strong>in</strong>(t)Jg i − cos(t)g j .<br />

From the local section i, we have<br />

i ∗ α ji = i ∗ ((π ∗ 1 ◦ s ∗ )(α ji )) = (π 1 ◦ i) ∗ (s ∗ ρα ji ),<br />

i ∗ β ji = i ∗ ((π ∗ 1 ◦ s ∗ )(β ji )) = (π 1 ◦ i) ∗ (s ∗ ρβ ji ).<br />

Thus, we have to study (s ∗ ρα ij )(ẋ kl ) and (s ∗ ρα ij )(ẋ l k ) (and the same for β ij). Then, we need<br />

(g l ◦ s ρ )(x ij ) s<strong>in</strong>ce<br />

Note that<br />

(s ∗ ρα ij )(ẋ kl ) = α ij (ds ρ (ẋ kl )) = 〈g j<br />

∣<br />

∣sρ(x<br />

kl (t)) , d(g i ◦ s ρ ) ∣ ∣<br />

sρ(x kl (t)) (ẋ kl)〉. (1.18)<br />

(g i ◦ s ρ )(x kl ) = i-th column of the matrix s ρ (x kl )<br />

and that s ρ (x kl ) ∈ G ′ is obta<strong>in</strong>ed from the parallel translation along the geodesic for O and<br />

with tangent vector g r+1 (t) <strong>in</strong> O. Hence, for curves x ij , x j i<br />

with i, j ≠ r + 1 we always take<br />

parallel translation along the same geodesic, when we apply s ρ .<br />

When we move g r+1 (t) we consider the parallel translation along different geodesics for each<br />

t. But, as curves x r+1,j , x j r+1 , x 1,r+1 just move the vector g r+1 (t) <strong>in</strong> a real plane generated by<br />

{g r+1 (0), g j (0)} or {g r+1 (0), Jg j (0)}, we have that g 0 (s ρ (x r+1,j )(t)) or g 0 (s ρ (x j r+1 (t))) describes<br />

a circle <strong>in</strong> CK n (ɛ) conta<strong>in</strong>ed <strong>in</strong> the plane generated by {g r+1 (0), g j (0)} (or {g r+1 (0), Jg j (0)}).<br />

From these remarks we have<br />

• (g 0 ◦ s ρ )(x kl ): po<strong>in</strong>t at a distance ρ from O <strong>in</strong> which we arrive along the geodesic with<br />

tangent vector g r+1 at O.<br />

⋄ x r+1,l , l > r + 1.<br />

⋄ x l r+1 , l ≥ r + 1.<br />

(g 0 ◦ s ρ )(x r+1,l ) = cos ɛ (ρ)O + s<strong>in</strong> ɛ (ρ)(cos(t)g r+1 + s<strong>in</strong>(t)g l ).<br />

(g 0 ◦ s ρ )(x l r+1) = cos ɛ (ρ)O + s<strong>in</strong> ɛ (ρ)(cos(t)g r+1 + s<strong>in</strong>(t)(Jg l )).<br />

,


1.5 Space of <strong>complex</strong> r-planes 23<br />

⋄ x 1,r+1 .<br />

(g 0 ◦ s ρ )(x 1,r+1 ) = cos ɛ (ρ)O + s<strong>in</strong> ɛ (ρ)(− s<strong>in</strong>(t)g 1 + cos(t)g r+1 )<br />

s<strong>in</strong>ce g 1 (t) = g 1 cos(t) + g r+1 s<strong>in</strong>(t) and g r+1 (t) is perpendicular to g 1 (t) for all t.<br />

Then, g r+1 (t) = −g 1 s<strong>in</strong>(t) + g r+1 cos(t).<br />

⋄ x r+1<br />

1 .<br />

⋄ x kl , k < l, k, l ≠ r + 1.<br />

(g 0 ◦ s ρ )(x r+1<br />

1 ) = cos ɛ (ρ)O + s<strong>in</strong> ɛ (ρ)(s<strong>in</strong>(t)(Jg 1 ) + cos(t)g r+1 ).<br />

(g 0 ◦ s ρ )(x kl ) = cos ɛ (ρ)O + s<strong>in</strong> ɛ (ρ)g r+1 .<br />

⋄ x l k , k ≤ l, k, l ≠ r + 1. (g 0 ◦ s ρ )(x l k ) = cos ɛ(ρ)O + s<strong>in</strong> ɛ (ρ)g r+1 .<br />

• (g r+1 ◦ s ρ )(x kl ):<br />

⋄ x r+1,l , l > r + 1.<br />

(g r+1 ◦ s ρ )(x r+1,l ) = s −1<br />

ρ (cos(t)g r+1 + s<strong>in</strong>(t)g l )<br />

but this parallel translation co<strong>in</strong>cides with the tangent vector to the geodesic at<br />

time ρ, that is,<br />

s −1<br />

ρ (cos(t)g r+1 +s<strong>in</strong>(t)g l ) = tangent vector to (cos ɛ (ρ)O+s<strong>in</strong> ɛ (ρ)(cos(t)g r+1 +s<strong>in</strong>(t)g l ))<br />

= s<strong>in</strong> ɛ (ρ)O + cos ɛ (ρ)(cos(t)g r+1 + s<strong>in</strong>(t)g l ).<br />

⋄ x l r+1 , l ≥ r + 1.<br />

⋄ x 1,r+1 .<br />

⋄ x r+1<br />

1 .<br />

⋄ x kl , k < l, k, l ≠ j.<br />

(g r+1 ◦ s ρ )(x l r+1) = s<strong>in</strong> ɛ (ρ)O + cos ɛ (ρ)(cos(t)g r+1 + s<strong>in</strong>(t)(Jg l )).<br />

(g r+1 ◦ s ρ )(x 1,r+1 ) = s<strong>in</strong> ɛ (ρ)O + cos ɛ (ρ)(− s<strong>in</strong>(t)g 1 + cos(t)g r+1 ).<br />

(g r+1 ◦ s ρ )(x r+1<br />

1 ) = s<strong>in</strong> ɛ (ρ)O + cos ɛ (ρ)(s<strong>in</strong>(t)(Jg 1 ) + cos(t)g r+1 ).<br />

(g r+1 ◦ s ρ )(x kl ) = s<strong>in</strong> ɛ (ρ)O + cos ɛ (ρ)g r+1 .<br />

⋄ x l k , k ≤ l, k, l ≠ j. (g r+1 ◦ s ρ )(x l k ) = s<strong>in</strong> ɛ(ρ)O + cos ɛ (ρ)g r+1 .<br />

• (g j ◦ s ρ )(x kl ), j > r + 1:<br />

⋄ x jl , l > j.<br />

(g j ◦ s ρ )(x jl ) = s −1<br />

ρ (cos(t)g j + s<strong>in</strong>(t)g l ).


24 Spaces of constant holomorphic curvature<br />

⋄ x l j , l ≥ j.<br />

⋄ x lj , l < j.<br />

⋄ x j l , l ≤ j.<br />

⋄ x kl , k < l, k, l ≠ j.<br />

⋄ x l k<br />

, k ≤ l, k, l ≠ j.<br />

(g<br />

(g j ◦ s ρ )(x l j) = s −1<br />

ρ (cos(t)g j + s<strong>in</strong>(t)(Jg l )).<br />

(g j ◦ s ρ )(x lj ) = s −1<br />

ρ (− s<strong>in</strong>(t)g l + cos(t)g j ).<br />

(g j ◦ s ρ )(x j l ) = s−1 ρ (s<strong>in</strong>(t)(Jg l ) + cos(t)g j ).<br />

(g j ◦ s ρ )(x kl ) = s −1<br />

ρ (g j ).<br />

j ◦ s ρ )(x l k ) = s−1 ρ (g j ).<br />

Now, we compute s ∗ ρα ij (and s ∗ ρβ ij ) <strong>in</strong> terms of α ij , β ij us<strong>in</strong>g (1.18) and evaluat<strong>in</strong>g s ∗ ρα ij<br />

(and s ∗ ρβ ij ) to each curve x kl . Do<strong>in</strong>g so, we obta<strong>in</strong><br />

(s ∗ ρα j1 ) g = −(α j1 ) g ′, j > r + 1.<br />

(s ∗ ρα r+1,1 ) g = − cos ɛ (ρ)(α r+1,1 ) g ′.<br />

(s ∗ ρβ j1 ) g = (β j1 ) g ′, j > r + 1.<br />

(s ∗ ρα r+1,1 ) g = − cos ɛ (ρ)(β r+1,1 ) g ′.<br />

F<strong>in</strong>ally, we have<br />

s ∗ ρ(<br />

∧<br />

j=r+1,...,n<br />

i=1,...,r<br />

α ji ∧ β j1 ) g = cos 2r<br />

ɛ (ρ)(<br />

∧<br />

j=r+1,...,n<br />

i=1,...,r<br />

α j1 ∧ β j1 ) g ′. (1.19)<br />

To get an expression for<br />

i ∗ (<br />

∧<br />

j=r+1,...,n<br />

i=1,...,r<br />

α ji ∧ β j1 )<br />

we use the diagram (1.17) and the computation <strong>in</strong> (1.19), so that<br />

i ∗ (<br />

∧<br />

j=r+1,...,n<br />

i=1,...,r<br />

α ji ∧ β ji ) g = i ∗ ◦ π ∗ 1 ◦ s ∗ ρ(<br />

∧<br />

j=r+1,...,n<br />

i=1,...,r<br />

= i ∗ ◦ π ∗ 1(cos 2r<br />

ɛ (ρ)(<br />

∧<br />

j=r+1,...,n<br />

i=1,...,r<br />

α ji ∧ β ji ) g<br />

α ji ∧ β ji ) ′ g)<br />

= cos 2r<br />

ɛ (ρ)π ∗ 3(dL r[0] ) = cos 2r<br />

ɛ (ρ)dL (n−r)[0]<br />

where we used the expression for the <strong>in</strong>variant density of <strong>complex</strong> r-planes through a po<strong>in</strong>t,<br />

cf. (1.20), and the duality between <strong>complex</strong> r-planes through a po<strong>in</strong>t and the <strong>complex</strong> (n − r)-<br />

planes through the same po<strong>in</strong>t.<br />

Hence, we get<br />

dL r = cos 2r<br />

ɛ (ρ)dx n−r dL (n−r)[O] .


1.5 Space of <strong>complex</strong> r-planes 25<br />

1.5.2 Density of <strong>complex</strong> r-planes conta<strong>in</strong><strong>in</strong>g a fixed <strong>complex</strong> q-plane<br />

We denote by L C r[q]<br />

the <strong>space</strong> of <strong>complex</strong> r-planes conta<strong>in</strong><strong>in</strong>g a fixed <strong>complex</strong> q-plane.<br />

We denote by H [q] := U ɛ (q) the isotropy group of a <strong>complex</strong> q-plane <strong>in</strong> CK n (ɛ) and by H r[q]<br />

the isotropy group of a <strong>complex</strong> r-plane conta<strong>in</strong><strong>in</strong>g the fixed <strong>complex</strong> q-plane. Note that H [q]<br />

acts transitively on L C r[q] .<br />

On the other hand, if we suppose that L 0 q is the fixed <strong>complex</strong> q-plane, then we can def<strong>in</strong>e<br />

the projection<br />

π : H [q] −→ H [q] /H r[q]<br />

g ↦→ L r[q] = gL 0 q.<br />

As the elements <strong>in</strong> H r[q] do not mix tangent vectors to the fixed <strong>complex</strong> q-plane with<br />

orthogonal vectors to this <strong>complex</strong> q-plane, we have<br />

⎧⎛<br />

⎞<br />

⎫<br />

⎨ A 0 0<br />

⎬<br />

H r[q]<br />

∼ = ⎝ 0 B 0 ⎠ , A ∈ U ɛ (q), B ∈ U(r − q), C ∈ U(n − r)<br />

⎩<br />

⎭ ,<br />

0 0 C<br />

π ∗ dL r[q] =<br />

∧<br />

q+1≤i≤r<br />

r+1≤j≤n<br />

α ij ∧ β ij . (1.20)<br />

(The <strong>forms</strong> vanish<strong>in</strong>g, with respect to the isometry group, <strong>in</strong> this case H [q] , are the ones<br />

<strong>in</strong>side the big box.)<br />

1.5.3 Density of <strong>complex</strong> q-planes conta<strong>in</strong>ed <strong>in</strong> a fixed <strong>complex</strong> r-plane<br />

Denote by L (r)<br />

q the <strong>space</strong> of <strong>complex</strong> q-planes conta<strong>in</strong>ed <strong>in</strong> a fixed <strong>complex</strong> r-plane. Let us<br />

fix a <strong>complex</strong> r-plane L r and a <strong>complex</strong> q-plane L (r)<br />

q conta<strong>in</strong>ed <strong>in</strong> L r . Consider the projection<br />

π : U ɛ (r) → U ɛ (r)/U ɛ (q) × U(r − q) where U ɛ (r) denotes the isometry group of the fixed<br />

<strong>complex</strong> r-plane and U ɛ (q) × U(r − q) the isotropy group of a <strong>complex</strong> q-plane conta<strong>in</strong>ed <strong>in</strong><br />

the <strong>complex</strong> r-plane. Then, as <strong>in</strong> the previous case we get<br />

π ∗ dL (r)<br />

q = ∧<br />

1≤i≤q<br />

q+1≤j≤r<br />

α ij ∧ β ij<br />

∧<br />

q+1≤j≤r<br />

α j0 ∧ β j0 =<br />

∧<br />

0≤i≤q<br />

q+1≤j≤r<br />

α ij ∧ β ij .<br />

1.5.4 Measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g a geodesic ball<br />

In order to obta<strong>in</strong> the value of this measure we use the expression for the <strong>in</strong>variant density<br />

of <strong>complex</strong> r-planes <strong>in</strong> (1.16) and the expression of the Jacobian of the map of chang<strong>in</strong>g to<br />

spherical coord<strong>in</strong>ates. This is given by (cf. [Gra73])<br />

cos ɛ (R) s<strong>in</strong>ɛ<br />

2n−1 (R)<br />

|ɛ| (n−1)/2 .<br />

Recall that cos ɛ and s<strong>in</strong> ɛ denote the generalized trigonometric functions def<strong>in</strong>ed at Notation<br />

1.2.2.<br />

We fix as a orig<strong>in</strong> of the spherical coord<strong>in</strong>ates the center of the geodesic ball. Then, us<strong>in</strong>g<br />

the expression <strong>in</strong> spherical coord<strong>in</strong>ates for the element volume of CK n−r (ɛ) (the orthogonal


26 Spaces of constant holomorphic curvature<br />

<strong>space</strong> to a <strong>complex</strong> r-plane <strong>in</strong>tersect<strong>in</strong>g the sphere), we obta<strong>in</strong><br />

∫<br />

∫ ∫<br />

m(L r ∈ L C r : B R ∩ L r ≠ ∅) = dL r =<br />

cos 2r<br />

ɛ (ρ)dx n−r ∧ dG C<br />

B R ∩L r≠∅<br />

G C n,n−r[O]<br />

n,n−r B R ∩L r<br />

∫<br />

= vol(GC n,n−r)<br />

|ɛ| (n−1)/2<br />

cos 2r+1<br />

ɛ (ρ) s<strong>in</strong> 2(n−r)−1<br />

ɛ (ρ)dρdS 2(n−r)−1<br />

∫ R<br />

S 2(n−r)−1 0<br />

= vol(GC n,n−r)<br />

|ɛ| (n−1)/2 vol(S2(n−r)−1 )<br />

= vol(GC n,n−r)<br />

|ɛ| (n−1)/2 vol(S2(n−r)−1 )<br />

= vol(GC n,n−r)<br />

|ɛ| (n−1)/2 vol(S2(n−r)−1 )<br />

= vol(GC n,n−r)<br />

|ɛ| (n−1)/2 vol(S2(n−r)−1 )<br />

∫ R<br />

0<br />

∫ R<br />

0<br />

cos 2r+1<br />

ɛ<br />

(ρ) s<strong>in</strong>ɛ<br />

2(n−r)−1 (ρ)dρ<br />

cos ɛ (ρ)(1 + s<strong>in</strong> 2 ɛ(ρ)) r s<strong>in</strong> 2(n−r)−1<br />

ɛ (ρ)<br />

r∑<br />

( ∫ r R<br />

i)<br />

i=0<br />

r∑<br />

i=0<br />

0<br />

cos ɛ (ρ) s<strong>in</strong>ɛ<br />

2(n−r+i)−1 (ρ)<br />

( 2(n−r+i)<br />

r s<strong>in</strong> ɛ (R)<br />

i)<br />

2(n − r + i) .<br />

At Chapter 4 we give a general expression for the measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g<br />

a regular doma<strong>in</strong>, <strong>in</strong> a way such that the previous expression can be <strong>in</strong>terpreted <strong>in</strong> terms of<br />

mean curvature <strong>in</strong>tegrals and other valuations def<strong>in</strong>ed at Chapter 2.<br />

1.5.5 Reproductive property of Quermass<strong>in</strong>tegrale<br />

At Chapter 3 we prove that the mean curvature <strong>in</strong>tegral does not satisfy a reproductive<br />

property (see Def<strong>in</strong>ition 3.4.1). In this section we prove that Quermass<strong>in</strong>tegrale do satisfy a<br />

reproductive property.<br />

Def<strong>in</strong>ition 1.5.9. Let Ω be a doma<strong>in</strong> <strong>in</strong> CK n (ɛ). For r ∈ {1, . . . , n − 1} we def<strong>in</strong>e<br />

W r (Ω) = (n − r) · O ∫<br />

r−1 · · · O 0<br />

χ(Ω ∩ L r )dL r<br />

n · O n−2 · · · O n−r−1<br />

where O i denotes the area of the sphere of radius <strong>in</strong> the standard Euclidean <strong>space</strong>. Moreover,<br />

we def<strong>in</strong>e<br />

W 0 (Ω) = vol(Ω) and W n (Ω) = O n−1<br />

n<br />

χ(Ω).<br />

Constants are chosen for analogy to the case of real <strong>space</strong> <strong>forms</strong>.<br />

Proposition 1.5.10. Let Ω be a doma<strong>in</strong> <strong>in</strong> CK n (ɛ). Then,<br />

∫<br />

W r (Ω) = c W r (Ω ∩ L q )dL q<br />

for 1 ≤ r ≤ q ≤ n − 1 and c is a constant depend<strong>in</strong>g only on n, r and q.<br />

L C q<br />

Proof. This proof is analogous to the one given by Santaló (cf. [San04]) to obta<strong>in</strong> the result<br />

<strong>in</strong> the Euclidean <strong>space</strong>.<br />

By def<strong>in</strong>ition, it is satisfied<br />

∫<br />

L C q<br />

W r (Ω ∩ L q )dL q = (q − r)O r−1 . . . O 0<br />

qO q−1 . . . O q−r−1<br />

L C r<br />

∫<br />

L C q<br />

∫<br />

Ω∩L r<br />

(q) ≠∅<br />

dL (q)<br />

r ∧ dL q .<br />

We express the densities dL (q)<br />

r , dL q , dL q[r] , dL r <strong>in</strong> terms of the <strong>forms</strong> ϕ ij def<strong>in</strong>ed at Lemma<br />

1.2.7 (we omit the absolute value for the densities),


1.5 Space of <strong>complex</strong> r-planes 27<br />

and<br />

Thus, the equality<br />

dL (q)<br />

r ∧ dL q = ∧<br />

dL q[r] ∧ dL r =<br />

holds.<br />

Apply<strong>in</strong>g it we get the result<br />

∫ ∫<br />

L C q<br />

Ω∩L (q)<br />

r ≠∅<br />

dL q =<br />

∧<br />

q


Chapter 2<br />

Introduction to valuations<br />

The notion of valuation <strong>in</strong> R n was <strong>in</strong>troduced by Blaschke <strong>in</strong> 1955 at [Bla55]. Recently,<br />

Alesker, among others, has extended this notion to differentiable manifolds. A survey about<br />

the development of valuations is given at [MS83] where some references are given.<br />

In this chapter we briefly <strong>in</strong>troduce the theory of valuations, focus<strong>in</strong>g on the concepts and<br />

results we shall use <strong>in</strong> the follow<strong>in</strong>g chapters.<br />

In the last section, we def<strong>in</strong>e some valuations <strong>in</strong> the <strong>space</strong>s of constant holomorphic curvature,<br />

generaliz<strong>in</strong>g the def<strong>in</strong>ition of some valuations <strong>in</strong> C n . We also give relations among the<br />

def<strong>in</strong>ed valuations.<br />

2.1 Def<strong>in</strong>ition and basic properties<br />

Let V be a vector <strong>space</strong> of real dimension n. We denote by K(V ) the set of non-empty compact<br />

convex doma<strong>in</strong>s <strong>in</strong> V .<br />

Def<strong>in</strong>ition 2.1.1. A functional φ : K(V ) → R is called a valuation if<br />

whenever A, B, A ∪ B ∈ K(V ).<br />

φ(A ∪ B) = φ(A) + φ(B) − φ(A ∩ B)<br />

Remark 2.1.2. The extension theorem of Groemer states that every valuation extends uniquely<br />

to the set of f<strong>in</strong>ite union of convex set.<br />

First examples of valuations <strong>in</strong> R n are the volume of a convex doma<strong>in</strong>, the area of its<br />

boundary and its Euler characteristic. Intr<strong>in</strong>sic volumes, def<strong>in</strong>ed by the coefficients of the<br />

polynomial obta<strong>in</strong>ed <strong>in</strong> the Ste<strong>in</strong>er’s formula, are also valuations.<br />

Consider <strong>in</strong> R n a convex doma<strong>in</strong> Ω and denote by Ω r the parallel doma<strong>in</strong> at a distance r<br />

from Ω. Recall that the parallel doma<strong>in</strong> is constituted by all po<strong>in</strong>ts at a distance less or equal<br />

than r.<br />

The Ste<strong>in</strong>er formula relates the volume of the parallel doma<strong>in</strong> with the volume and some<br />

other functionals of the <strong>in</strong>itial doma<strong>in</strong>.<br />

Proposition 2.1.3 (Ste<strong>in</strong>er’s formula). Let Ω ⊂ R n be a compact doma<strong>in</strong> and Ω r the parallel<br />

doma<strong>in</strong> at distance r. Then, the volume of Ω r can be expressed as a polynomial <strong>in</strong> r and<br />

its coefficients are multiples of the valuations V i : K(V ) → R, i ∈ {0, . . . , n}, called <strong>in</strong>tr<strong>in</strong>sic<br />

volumes. The explicit expression is<br />

n∑<br />

vol(Ω r ) = r n−i ω n−i V i (Ω) (2.1)<br />

i=0<br />

where ω n−i denotes the volume of the (n − i)-dimensional ball of radius 1 <strong>in</strong> the Euclidean<br />

<strong>space</strong>.<br />

29


30 Introduction to valuations<br />

Proof. (of the fact that V i are valuations.)<br />

Let A, B, A ∪ B ∈ K(V ). Then, it is satisfied (A ∩ B) r = A r ∩ B r and (A ∪ B) r = A r ∪ B r .<br />

Thus,<br />

vol((A ∪ B) r ) = vol(A r ) + vol(B r ) − vol(A r ∩ B r ) = vol(A r ) + vol(B r ) − vol((A ∩ B) r )<br />

∀r.<br />

Apply<strong>in</strong>g (2.1) we deduce that the <strong>in</strong>tr<strong>in</strong>sic volumes are valuations.<br />

Some particular cases of <strong>in</strong>tr<strong>in</strong>sic volumes are<br />

• V n (Ω) = vol(Ω),<br />

• V n−1 (Ω) = vol(∂Ω)/2,<br />

• V 0 (Ω) = χ(Ω).<br />

Note that the Ste<strong>in</strong>er formula has sense for any convex doma<strong>in</strong>, without any assumption<br />

on the regularity of the boundary. In some cases it is <strong>in</strong>terest<strong>in</strong>g to consider convex doma<strong>in</strong>s<br />

such that its boundary is an oriented hypersurface of class C 2 . Apply<strong>in</strong>g the previous formula<br />

<strong>in</strong> this case we obta<strong>in</strong> the so-called mean curvature <strong>in</strong>tegrals.<br />

Def<strong>in</strong>ition 2.1.4. Let S be a hypersurface of class C 2 <strong>in</strong> a Riemannian manifold M of dimension<br />

n. If x ∈ S, we denote the second fundamental form of S at x by II x . We def<strong>in</strong>e the i-th<br />

mean curvature <strong>in</strong>tegral of S as<br />

( ) n − 1 −1 ∫<br />

M i (S) =<br />

i<br />

S<br />

σ i (II x )dx<br />

where σ i (II x ) denotes the r-th symmetric elementary function of the second fundamental form<br />

II x .<br />

Then, the Ste<strong>in</strong>er formula <strong>in</strong> R n is<br />

n−1<br />

(<br />

∑ n<br />

)<br />

vol(Ω r ) = vol(Ω) + r n−i i<br />

n M n−i−1(∂Ω).<br />

i=0<br />

Sometimes, it is def<strong>in</strong>ed M −1 (∂Ω) := vol(Ω).<br />

The relation among the <strong>in</strong>tr<strong>in</strong>sic volumes and the mean curvature <strong>in</strong>tegrals, for convex<br />

doma<strong>in</strong>s with boundary of class C 2 , is<br />

( n<br />

i)<br />

V i (Ω) = M n−i−1 (∂Ω).<br />

nω n−i<br />

Def<strong>in</strong>ition 2.1.5. A valuation φ is cont<strong>in</strong>uous if it is cont<strong>in</strong>uous with respect to the Hausdorff<br />

metric.<br />

Recall that the Hausdorff distance between two compact sets A, B is given by<br />

d Haus (A, B) = max{sup <strong>in</strong>f {d(a, b)}, sup <strong>in</strong>f {d(a, b)}}<br />

a∈A b∈B<br />

b∈B a∈A<br />

where d(a, b) is the distance def<strong>in</strong>ed <strong>in</strong> the ambient <strong>space</strong> for A and B.<br />

Example 2.1.6. The <strong>in</strong>tr<strong>in</strong>sic volumes are cont<strong>in</strong>uous valuations. Anyway, there are some<br />

<strong>in</strong>terest<strong>in</strong>g examples of non-cont<strong>in</strong>uous valuations <strong>in</strong> R n . For example, the aff<strong>in</strong>e surface area<br />

is a valuation <strong>in</strong> the Euclidean <strong>space</strong>, but it is not cont<strong>in</strong>uous (cf. [KR97]). The aff<strong>in</strong>e surface<br />

area of a convex doma<strong>in</strong> Ω ⊂ R n is def<strong>in</strong>ed as the <strong>in</strong>tegral of the (n + 1)-th root of the


2.1 Def<strong>in</strong>ition and basic properties 31<br />

generalized Gauss curvature (cf. [Sch93] notes of Sections 1.5 and 2.5) of the boundary of the<br />

doma<strong>in</strong> with respect to the (n − 1)-dimensional Hausdorff measure of the boundary<br />

∫<br />

n+1<br />

AS(Ω) =<br />

√ K x dx.<br />

∂Ω<br />

One of the most important property of this valuation is that it is <strong>in</strong>variant under translations<br />

and l<strong>in</strong>ear transformations with determ<strong>in</strong>ant 1.<br />

Def<strong>in</strong>ition 2.1.7. Given a (2n − 1)-form ω def<strong>in</strong>ed on S(V ), and a smooth measure, η we<br />

consider, for each regular doma<strong>in</strong> Ω,<br />

∫ ∫<br />

η + ω<br />

N(Ω)<br />

Ω<br />

where N(Ω) denotes the normal bundle of the boundary of the doma<strong>in</strong>. The obta<strong>in</strong>ed functional<br />

is called smooth valuation.<br />

Def<strong>in</strong>ition 2.1.8. Let Ω ∈ K(V ). A valuation φ : K(V ) → R is called<br />

• translation <strong>in</strong>variant if<br />

φ(ψΩ) = φ(Ω)<br />

for every ψ translation of the vector <strong>space</strong> V ;<br />

• <strong>in</strong>variant with respect to a group G act<strong>in</strong>g on V if<br />

for every g ∈ G;<br />

• homogeneous of degree k if<br />

φ(gΩ) = φ(Ω)<br />

φ(λΩ) = λ k φ(Ω) for every λ > 0, k ∈ R;<br />

• even (resp. odd) if<br />

with ɛ even (resp. odd);<br />

φ(−1 · Ω) = (−1) ɛ φ(Ω)<br />

• monotone if<br />

φ(Ω 1 ) ≥ φ(Ω 2 ) for every Ω 1 , Ω 2 ∈ K(V ) and Ω 1 ⊃ Ω 2 .<br />

The <strong>space</strong> of cont<strong>in</strong>uous <strong>in</strong>variant translation valuations is denoted by Val(V ), the sub<strong>space</strong><br />

of Val(V ) of the homogeneous valuations of degree k by Val k (V ) and the sub<strong>space</strong> of Val(V )<br />

of even valuations (resp. odd valuations) by Val + (V ) (resp. Val − (V )).<br />

Example 2.1.9. The <strong>in</strong>tr<strong>in</strong>sic volume V i is a cont<strong>in</strong>uous <strong>in</strong>variant translation valuation homogeneous<br />

of degree i.<br />

Remark 2.1.10. The <strong>space</strong> Val(V ) has structure of <strong>in</strong>f<strong>in</strong>ite dimensional vector <strong>space</strong>.<br />

The follow<strong>in</strong>g result of P. McMullen [McM77] gives a decomposition of the <strong>space</strong> of valuations<br />

depend<strong>in</strong>g on the degree, and another depend<strong>in</strong>g on the parity<br />

Theorem 2.1.11 ([McM77]). Let n = dim V . Then,<br />

Val(V ) =<br />

n⊕<br />

Val i (V ) and Val(V ) = Val + (V ) ⊕ Val − (V ).<br />

i=0


32 Introduction to valuations<br />

The l<strong>in</strong>ear group GL(V ) of <strong>in</strong>vertible l<strong>in</strong>ear transformations of V acts transitively on Val(V )<br />

(gφ)(Ω) = φ(g −1 (Ω)) for g ∈ GL(V ), φ ∈ Val(V ), Ω ∈ K(V ).<br />

This action is cont<strong>in</strong>uous and preserves the homogeneous degree of the valuation.<br />

Theorem 2.1.12 (Irreducibility Theorem). Let V be an n-dimensional vector <strong>space</strong>. The<br />

natural representation of GL(V ) at Val + i (V ) and Val− i<br />

(V ) is irreducible for any i ∈ {0, . . . , n}.<br />

That is, there is no proper closed GL(V )-<strong>in</strong>variant sub<strong>space</strong>.<br />

From this theorem, it can be proved the follow<strong>in</strong>g result which relates cont<strong>in</strong>uous valuations<br />

with smooth valuations.<br />

Theorem 2.1.13 ([Ale01]). In a vector <strong>space</strong> V , the smooth translation <strong>in</strong>variant valuations<br />

are dense <strong>in</strong> the <strong>space</strong> of cont<strong>in</strong>uous translation <strong>in</strong>variant valuations.<br />

If V has an Euclidean metric, then every group G subgroup of the orthogonal group, acts on<br />

Val(V ) and it has sense to consider the <strong>space</strong> Val G (V ) ⊂ Val(V ), i.e. the <strong>space</strong> of G-<strong>in</strong>variant<br />

valuations under the action of the group G ⋉ V . The follow<strong>in</strong>g result by Alesker gives the<br />

necessary and sufficient condition to be this <strong>space</strong> of f<strong>in</strong>ite dimension.<br />

Corollary 2.1.14 ([Ale07a] Proposition 2.6). The <strong>space</strong> Val G (V ) is f<strong>in</strong>ite dimensional if and<br />

only of G acts transitively on the unit sphere of V .<br />

2.2 Hadwiger Theorem<br />

In 1957 Hadwiger proved the follow<strong>in</strong>g result concern<strong>in</strong>g valuations.<br />

Theorem 2.2.1 ([Had57]). The dimension of the <strong>space</strong> of cont<strong>in</strong>uous translation and O(n)-<br />

<strong>in</strong>variant valuations is<br />

dim Val O(n) (R n ) = n + 1<br />

and a basis of this <strong>space</strong> is given by<br />

where V i denotes the i-th <strong>in</strong>tr<strong>in</strong>sic volume.<br />

V 0 , V 1 , . . . , V n−1 , V n<br />

Remark 2.2.2. From this theorem it follows that <strong>in</strong> R n the sub<strong>space</strong> of valuations of homogeneous<br />

degree k ∈ {0, . . . , n} is of dimension 1.<br />

Last remark allows us to prove, <strong>in</strong> an easy way, some of the classical results of <strong>in</strong>tegral<br />

<strong>geometry</strong> (<strong>in</strong> R n ), such as reproductive or the k<strong>in</strong>ematic formulas.<br />

Example 2.2.3. • Crofton formula. Let Ω ⊂ R n be a compact convex doma<strong>in</strong>, and let L r<br />

be the <strong>space</strong> of all planes of dimension r with dL r the (unique up to a constant factor)<br />

<strong>in</strong>variant density. The measure of the set of planes a convex body <strong>in</strong> R n can be expressed<br />

<strong>in</strong> terms of the <strong>in</strong>tr<strong>in</strong>sic volumes as<br />

∫<br />

χ(Ω ∩ L r )dL r = cV n−r (Ω).<br />

L r<br />

This expression is obta<strong>in</strong>ed from the fact that the <strong>in</strong>tegral <strong>in</strong> the left hand side is a<br />

valuation of homogeneous degree (n − r).<br />

At Chapter 4, we study the expression of the measure of <strong>complex</strong> planes meet<strong>in</strong>g a<br />

doma<strong>in</strong> <strong>in</strong> CK n (ɛ), and at Chapter 5, the measure of totally real planes of dimension n<br />

meet<strong>in</strong>g a doma<strong>in</strong>.


2.2 Hadwiger Theorem 33<br />

• Reproductive property of the <strong>in</strong>tr<strong>in</strong>sic volumes. If Ω is a compact convex doma<strong>in</strong>, the<br />

reproductive formula for the <strong>in</strong>tr<strong>in</strong>sic volumes <strong>in</strong> R n is given by<br />

∫<br />

V (r)<br />

i<br />

L r<br />

(∂Ω ∩ L r )dL r = cV i (∂Ω)<br />

where L r denotes the <strong>space</strong> of r-planes <strong>in</strong> R n and c is a constant depend<strong>in</strong>g only on n,<br />

r and i.<br />

Clearly, the <strong>in</strong>tegral <strong>in</strong> the left hand side is a cont<strong>in</strong>uous valuation of homogeneous<br />

degree i. By the Hadwiger Theorem the equality follows directly, s<strong>in</strong>ce the i-th <strong>in</strong>tr<strong>in</strong>sic<br />

volume is a homogeneous valuation of this degree and the dimension of the <strong>space</strong> of<br />

homogeneous valuations of degree i is 1. In order to compute the value of c it is used<br />

the so-called template method, which consists on evaluat<strong>in</strong>g each side of the equation <strong>in</strong><br />

an easy doma<strong>in</strong> (for <strong>in</strong>stance, a sphere) and then compute the value of c. In [San04] it is<br />

given a way to prove this reproductive formula, but us<strong>in</strong>g the mean curvature <strong>in</strong>tegrals.<br />

The value of c it is also computed.<br />

• K<strong>in</strong>ematic formula. Although <strong>in</strong> this work we do not study k<strong>in</strong>ematic formulas, we would<br />

like to give, for completion, the classical k<strong>in</strong>ematic formula of Blaschke-Santaló.<br />

One of the problems of study of the <strong>in</strong>tegral <strong>geometry</strong> consists on measur<strong>in</strong>g the movements<br />

of R n which takes one convex doma<strong>in</strong> to another fixed one. In R n , if we denote<br />

by O(n) the movements group of the <strong>space</strong>, we obta<strong>in</strong> the classical k<strong>in</strong>ematic formula<br />

∫<br />

O(n)<br />

χ(Ω 1 ∩ gΩ 2 )dg =<br />

n∑<br />

c n,i V i (Ω 1 )V n−i (Ω 2 ). (2.2)<br />

This formula can be proved from apply<strong>in</strong>g twice the Hadwiger Theorem, and then,<br />

apply<strong>in</strong>g the obta<strong>in</strong>ed formula to spheres of different radius.<br />

Note that the <strong>in</strong>tegral on the left hand side is a functional on the first convex doma<strong>in</strong><br />

Ω 1 , but also on the second convex doma<strong>in</strong> Ω 2 . As a functional on the second convex<br />

doma<strong>in</strong>, from the Hadwiger Theorem, we have<br />

∫<br />

n∑<br />

χ(Ω 1 ∩ gΩ 2 )dg = c i (Ω 1 )V i (Ω 2 )<br />

O(n)<br />

where the coefficients c i (Ω 1 ) depend on Ω 1 . But, the <strong>in</strong>tegral under consideration is also<br />

a valuation with respect to Ω 1 , thus, the coefficients c i (Ω 1 ) are valuations and, aga<strong>in</strong> by<br />

Hadwiger Theorem, we obta<strong>in</strong> that it is satisfied<br />

∫<br />

n∑ n∑<br />

χ(Ω 1 ∩ gΩ 2 )dg = c ij V j (Ω 1 )V i (Ω 2 ).<br />

O(n)<br />

i=0<br />

i=0<br />

i=0 j=0<br />

To obta<strong>in</strong> the expression (2.2), first, note that the desired expression have to be symmetric<br />

with respect to Ω 1 and Ω 2 , hence, c ij = c ji . To prove that most of the constants<br />

c ij vanishes we use the template method, i.e. we apply the equality for a sphere of radius<br />

r and for a sphere of radius R.<br />

Us<strong>in</strong>g the <strong>in</strong>variance with respect to the rotations of a sphere we have<br />

∫<br />

∫ ∫<br />

χ(B r ∩ gB R )dg = χ(B r ∩ (φB R + v))dvdφ (2.3)<br />

O(n)<br />

O(n) R n ∫<br />

= vol(O(n)) χ(B r ∩ (B R + v))dv = vol(O(n))(r + R) n ω n .<br />

R n


34 Introduction to valuations<br />

On the other hand, as the <strong>in</strong>tr<strong>in</strong>sic volume V i is a homogeneous valuation of degree i we<br />

get<br />

n∑ n∑<br />

n∑<br />

c ij V j (B r )V i (B R ) = c ij r j R i V j (B 1 )V i (B 1 ). (2.4)<br />

i=0 j=0<br />

Thus, <strong>in</strong> both cases (2.3) and (2.4), we obta<strong>in</strong>ed a polynomial on r and R. Compar<strong>in</strong>g<br />

the coefficients we get i + j = n <strong>in</strong> the last summation <strong>in</strong> (2.4).<br />

To get the explicit value for the constant, it rema<strong>in</strong>s only to use the expression of the<br />

<strong>in</strong>tr<strong>in</strong>sic volume of sphere of any radius. As the parallel doma<strong>in</strong> at distance r of a sphere<br />

of radius R is a sphere of radius r + R, we can easily compute its <strong>in</strong>tr<strong>in</strong>sic volume us<strong>in</strong>g<br />

the Ste<strong>in</strong>er formula, and we get<br />

2.3 Alesker Theorem<br />

i,j=0<br />

V i (B r ) = r i ( n<br />

i<br />

)<br />

ωn<br />

ω n−i<br />

.<br />

Recently, Alesker gave an analogous theorem of Hadwiger Theorem for the Hermitian standard<br />

<strong>space</strong> V = C n , with isometry group IU(n) = C n ⋊U(n). As the isometry group of C n is smaller<br />

than the isometry group of R 2n it may happen that some non-<strong>in</strong>variant valuations under the<br />

isometry group of R 2n is <strong>in</strong>variant under the isometry group of C n , and this occurs.<br />

Theorem 2.3.1 ([Ale03] Theorem 2.1.1). Let Val U(n) (C n ) be the <strong>space</strong> of cont<strong>in</strong>uous translation<br />

and U(n)-<strong>in</strong>variant valuations <strong>in</strong> C n . Then,<br />

( ) n + 2<br />

dim Val U(n) (C n ) = ,<br />

2<br />

and the dimension of the sub<strong>space</strong> of degree k homogeneous valuations is<br />

m<strong>in</strong>{k, 2n − k}<br />

2<br />

+ 1.<br />

Alesker [Ale03] also gave two bases of cont<strong>in</strong>uous isometry <strong>in</strong>variant valuations on C n . One<br />

of these bases is def<strong>in</strong>ed as the <strong>in</strong>tegral of the projection volume. That is, let Ω ⊂ C n be a<br />

convex doma<strong>in</strong> and k, l <strong>in</strong>tegers such that 0 ≤ k ≤ 2l ≤ 2n, then<br />

∫<br />

C k,l (Ω) :=<br />

G C n,l<br />

V k (Pr Ll (Ω))dL l<br />

where G C n,l denotes the <strong>space</strong> of <strong>complex</strong> l-planes <strong>in</strong> Cn through the orig<strong>in</strong> (see Section 1.5),<br />

Pr Ll (Ω) denotes the orthogonal projection of Ω at L l and V k the k-th <strong>in</strong>tr<strong>in</strong>sic volume. Valuations<br />

{C k,l } with 0 ≤ k ≤ 2l ≤ 2n def<strong>in</strong>e a basis of Val U(n) (C n ). The <strong>in</strong>dex k co<strong>in</strong>cides with<br />

the homogeneous degree of the valuation.<br />

The other basis of Val U(n) (C n ) is {U k,p } with k, p <strong>in</strong>tegers such that 0 ≤ 2p ≤ k ≤ 2n with<br />

∫<br />

U k,p (Ω) :=<br />

L C n−p<br />

V k−2p (Ω ∩ L n−p )dL n−p (2.5)<br />

where L C n−p denotes the <strong>space</strong> of <strong>complex</strong> aff<strong>in</strong>e (n − p)-planes of C n (see Section 1.5), and<br />

V k−2p the (k − 2p)-th <strong>in</strong>tr<strong>in</strong>sic volume. The <strong>in</strong>dex k co<strong>in</strong>cides with the homogeneous degree of<br />

the valuation.


2.3 Alesker Theorem 35<br />

Example 2.3.2. We describe explicitily the elements of {U k,p } <strong>in</strong> C 3 . In C 3 there are ( 5<br />

2)<br />

= 10<br />

l<strong>in</strong>early <strong>in</strong>dependent valuations. For each degree k we have as much valuations as l<strong>in</strong>early<br />

<strong>in</strong>dependent <strong>in</strong>tegers p such that<br />

0 ≤ p ≤<br />

m<strong>in</strong>{k, 2n − k}<br />

.<br />

2<br />

Suppose that Ω is a convex doma<strong>in</strong> <strong>in</strong> C 3 . Then, a basis of Val U(3) (C 3 ) is given by the follow<strong>in</strong>g<br />

valuations.<br />

k = 0: There is only one value of p, p = 0 and<br />

k = 1: There is only one value of p, p = 0 and<br />

k = 2: There are two values of p, p = 0, 1.<br />

If p = 0 then<br />

U 0,0 (Ω) = V 0 (Ω) = χ(Ω).<br />

U 1,0 (Ω) = V 1 (Ω).<br />

U 2,0 (Ω) = V 2 (Ω).<br />

If p = 1 then<br />

∫<br />

U 2,1 (Ω) = V 0 (Ω ∩ L 2 )dL 2 .<br />

L 2<br />

k = 3: There are two values of p, p = 0, 1.<br />

If p = 0 then<br />

U 3,0 (Ω) = V 3 (Ω).<br />

If p = 1 then<br />

∫<br />

U 3,1 (Ω) = V 1 (Ω ∩ L 2 )dL 2 .<br />

L 2<br />

k = 4: There are two values of p, p = 0, 1.<br />

If p = 0 then<br />

U 4,0 (Ω) = V 4 (Ω).<br />

If p = 1 then<br />

∫<br />

U 4,1 (Ω) = V 2 (Ω ∩ L 2 )dL 2 .<br />

L 2<br />

k = 5: There is only one value of p, p = 0 and<br />

U 5,0 (Ω) = V 5 (Ω) = 1 2 vol(∂Ω).<br />

k = 6: There is only one value of p, p = 0 and<br />

U 6,0 (Ω) = V 6 (Ω) = vol(Ω).<br />

Note that we obta<strong>in</strong>ed all <strong>in</strong>tr<strong>in</strong>sic volumes V j (K), j ∈ {0, . . . , 6}, <strong>in</strong> the same way as <strong>in</strong><br />

R 6 , but at C 3 appear three new l<strong>in</strong>ear <strong>in</strong>dependent valuations.


36 Introduction to valuations<br />

In the same way as <strong>in</strong> R n , hav<strong>in</strong>g a basis for Val U(n) (C n ) allows us to establish a k<strong>in</strong>ematic<br />

formula. The difference is that, now, does not seem possible to f<strong>in</strong>d the constants us<strong>in</strong>g the<br />

template method. Alesker, <strong>in</strong> the same paper [Ale03] establish the follow<strong>in</strong>g result.<br />

Theorem 2.3.3 ([Ale03] Theorem 3.1.1). Let Ω 1 , Ω 2 ⊂ C n be doma<strong>in</strong>s with piecewise smooth<br />

boundaries such that for any U ∈ IU(n) the <strong>in</strong>tersection Ω 1 ∩ U(Ω 2 ) has a f<strong>in</strong>ite number of<br />

components. Then,<br />

∫<br />

χ(Ω 1 ∩ U(Ω 2 ))dU =<br />

∑ ∑<br />

κ(k 1 , k 2 , p 1 , p 2 )U k1 ,p 1<br />

(Ω 1 )U k2 ,p 2<br />

(Ω 2 ),<br />

U∈IU(n)<br />

p 1 ,p 2<br />

k 1 +k 2 =2n<br />

where the <strong>in</strong>dex of the <strong>in</strong>ner summation runs over 0 ≤ p i ≤ k i /2, i = 1, 2, and κ(k 1 , k 2 , p 1 , p 2 )<br />

are constants depend<strong>in</strong>g only on n, k 1 , k 2 , p 1 , p 2 .<br />

Theorem 2.3.4 ([Ale03] Theorem 3.1.2). Let Ω ⊂ C n be a doma<strong>in</strong> with piecewise smooth<br />

boundary and 0 < q < n, 0 < 2p < k < 2q. Then,<br />

∫<br />

L C r<br />

U k,p (Ω ∩ L r )dL r =<br />

[k/2]+n−q<br />

∑<br />

p=0<br />

where γ p are constants depend<strong>in</strong>g only on n, q, and p.<br />

γ p · U k+2(n−q),p (Ω),<br />

Theorem 2.3.5 ([Ale03] Theorem 3.1.3). Let Ω ⊂ C n be a doma<strong>in</strong> with piecewise smooth<br />

boundary. Then,<br />

∫<br />

[n/2]<br />

∑<br />

χ(Ω ∩ L n )dL n = β p · U n,p (Ω),<br />

L R n<br />

where L R n denotes the <strong>space</strong> of Lagrangian planes (i.e. totally real planes of dimension n) <strong>in</strong><br />

C n and β p are constants depend<strong>in</strong>g only on n and p.<br />

The constants for Theorem 2.3.3 are given by Bernig-Fu at [BF08]. These constants were<br />

computed us<strong>in</strong>g <strong>in</strong>direct methods and others bases of valuations <strong>in</strong> C n . In this work, we give<br />

the constant, <strong>in</strong> some cases, for Theorems 2.3.4 and 2.3.5.<br />

In [BF08] are given some other bases for Val U(n) (C n ). In Chapters 4 and 5, we use a basis<br />

def<strong>in</strong>ed <strong>in</strong> [BF08] and we extend it to all <strong>complex</strong> <strong>space</strong> <strong>forms</strong> CK n (ɛ). The def<strong>in</strong>ition of theses<br />

valuations and its extension <strong>in</strong> CK n (ɛ) is given at Section 2.4.2.<br />

F<strong>in</strong>ally, we note that Proposition 2.1.14 gives another way to generalize the theory of valuations<br />

<strong>in</strong> vector <strong>space</strong>s. From this proposition, we have that for any group act<strong>in</strong>g transitively<br />

on the sphere can be stated a Hadwiger type theorem, i.e. the <strong>space</strong> of cont<strong>in</strong>uous translation<br />

<strong>in</strong>variant valuations has f<strong>in</strong>ite dimension, thus, it has sense to compute its dimension and give<br />

a basis. An expression for the k<strong>in</strong>ematic formula can also be given.<br />

In this section, we recalled the case <strong>in</strong> which the act<strong>in</strong>g group is U(n) and befors we studied<br />

SO(n), but there are some known result for other groups.<br />

The groups act<strong>in</strong>g over a sphere are classified and they are (cf. [Bor49], [Bor50], [MS43])<br />

six <strong>in</strong>f<strong>in</strong>ite series<br />

and three exceptional groups<br />

p=0<br />

SO(n), U(n), SU(n), Sp(n), Sp(n) · U(1), Sp(n) · Sp(1)<br />

G 2 , Sp<strong>in</strong>(7), Sp<strong>in</strong>(9).<br />

Alesker and Bernig obta<strong>in</strong>ed a Hadwiger type theorem, a k<strong>in</strong>ematic formula (and the<br />

algebraic structure) for some of these groups (cf. [Ale04] for G = SU(2), [Ber08a] for G =<br />

SU(n) and [Ber08b] for G = G 2 and G = Sp<strong>in</strong>(7)).


2.4 Valuations on <strong>complex</strong> <strong>space</strong> <strong>forms</strong> 37<br />

2.4 Valuations on <strong>complex</strong> <strong>space</strong> <strong>forms</strong><br />

In this section we def<strong>in</strong>e some valuations on CK n (ɛ) and we give relations among them.<br />

2.4.1 Smooth valuations on manifolds<br />

The notion of smooth valuation <strong>in</strong> a differentiable manifold was recently studied (cf. [Ale06a],<br />

[Ale06b], [AF08], [Ale07b], [AB08]). First, a def<strong>in</strong>ition of smooth valuation on a manifold was<br />

given, and then it was proved that some important properties, which we do not study <strong>in</strong> this<br />

work, of smooth valuations, such as the duality property, still hold.<br />

Def<strong>in</strong>ition 2.1.7 can be extended for differentiable manifolds.<br />

Def<strong>in</strong>ition 2.4.1. Let M be a differentiable manifold and Ω a compact submanifold. Given<br />

a (2n − 1)-form ω <strong>in</strong> S(M), and a smooth measure η, we consider for any Ω<br />

∫ ∫<br />

η + ω,<br />

Ω<br />

where N(Ω) denotes the normal cycle (cf. [Ale07a]). The obta<strong>in</strong>ed functional is called smooth<br />

valuation.<br />

Remark 2.4.2. A more general def<strong>in</strong>ition analogue to Def<strong>in</strong>iton 2.1.1 appears <strong>in</strong> [Ale06b], and<br />

it is called f<strong>in</strong>itely additive measure.<br />

In <strong>space</strong>s of constant sectional curvatures, <strong>in</strong>variant smooth valuations are well-known.<br />

These <strong>space</strong>s have the same isotropy group of a po<strong>in</strong>t as a po<strong>in</strong>t <strong>in</strong> R n , and from the po<strong>in</strong>t of<br />

view of a homogeneous <strong>space</strong>s they can be studied <strong>in</strong> an analogous way. Despite this fact, a<br />

Hadwiger type theorem for cont<strong>in</strong>uous valuations (and not only for smooth valuations) it is<br />

not known, i.e. it is not known a basis of cont<strong>in</strong>uous translation <strong>in</strong>variant valuations <strong>in</strong>variant<br />

also for the isometry group of the <strong>space</strong>. The dimension of this <strong>space</strong> of valuations it is not<br />

known an analogous result to Theorem 2.1.13.<br />

Anyway, a big amount of the results <strong>in</strong> <strong>in</strong>tegral <strong>geometry</strong> are known <strong>in</strong> these <strong>space</strong>s. For<br />

<strong>in</strong>stance, Santaló [San04, page 309] proved that a reproductive formula holds for any real<br />

<strong>space</strong> form and also obta<strong>in</strong>ed an expression for the measure of totally geodesic planes meet<strong>in</strong>g<br />

a convex doma<strong>in</strong>.<br />

In view of these results of Santaló and the knowledge of a basis of cont<strong>in</strong>uous <strong>in</strong>variant<br />

valuations on C n , the aim of this work is to study the classical formulas <strong>in</strong> <strong>in</strong>tegral <strong>geometry</strong><br />

described <strong>in</strong> the last paragraph <strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong>, i.e. <strong>in</strong> the standard Hermitian <strong>space</strong>,<br />

and <strong>in</strong> the <strong>complex</strong> projective and hyperbolic <strong>space</strong>.<br />

2.4.2 Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes<br />

Bernig and Fu [BF08] def<strong>in</strong>ed the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes <strong>in</strong> C n . In this section we recall<br />

this def<strong>in</strong>ition and its extension to CK n (ɛ) follow<strong>in</strong>g [Par02].<br />

Bernig and Fu at [BF08, page 14] def<strong>in</strong>ed <strong>in</strong> T C n , the follow<strong>in</strong>g <strong>in</strong>variant 1-<strong>forms</strong> α, β<br />

and γ and the <strong>in</strong>variant 2-<strong>forms</strong> θ 0 , θ 1 , θ 2 and θ s . Let (z 1 , . . . , z n , ζ 1 , . . . , ζ n ) be the canonical<br />

coord<strong>in</strong>ates of T C n ≃ C n × C n with z i = x i + √ −1y i and ζ i = ξ i + √ −1η i . Then,<br />

N(Ω)<br />

θ 0 := ∑ n<br />

i=1 dξ i ∧ dη i , θ 1 := ∑ n<br />

i=1 (dx i ∧ dη i − dy i ∧ dξ i ) ,<br />

θ 2 := ∑ n<br />

i=1 dx i ∧ dy i , θ s := ∑ n<br />

i=1 (dx i ∧ dξ i + dy i ∧ dη i ) ,<br />

α := ∑ n<br />

i=1 ξ idx i + η i dy i , β := ∑ n<br />

i=1 ξ idy i − η i dx i ,<br />

γ := ∑ n<br />

i=1 ξ idη i − η i dξ i .


38 Introduction to valuations<br />

α is the contact form (see Remark 1.3.8) and θ s is the symplectic form of T C n . Recall that a 2-<br />

form ω def<strong>in</strong>ed <strong>in</strong> a manifold of dimension 2m is said symplectic if it is closed, non-degenerated<br />

and ω m ≠ 0.<br />

The previous <strong>forms</strong> are only def<strong>in</strong>ed for ɛ = 0 s<strong>in</strong>ce canonical coord<strong>in</strong>ates exist only at C n .<br />

But, we can express them <strong>in</strong> terms of a mov<strong>in</strong>g frame on S(C n ) (and, then extend them for<br />

any ɛ ∈ R). The expression <strong>in</strong> terms of a mov<strong>in</strong>g frame allows us to prove that these <strong>forms</strong> are<br />

well-def<strong>in</strong>ed, i.e. they do not depend on the chosen coord<strong>in</strong>ates. Let (x, v) ∈ S(CK n (ɛ)) and<br />

let {x; e 1 := v, Je 1 , . . . , e n , Je n } be a J-mov<strong>in</strong>g frame def<strong>in</strong>ed <strong>in</strong> a neighborhood of x. Then<br />

θ 0 =<br />

θ 1 =<br />

θ 2 =<br />

n∑<br />

α 1i ∧ β 1i ,<br />

i=2<br />

n∑<br />

(α i ∧ β 1i − β i ∧ α 1i ), (2.6)<br />

i=2<br />

n∑<br />

α i ∧ β i ,<br />

i=2<br />

where α i , β i , α ij , β ij are the <strong>forms</strong> given at (1.12) but <strong>in</strong>terpreted as <strong>forms</strong> <strong>in</strong> S(CK n (ɛ)).<br />

Remark 2.4.3. From the expression of θ 0 , θ 1 and θ 2 <strong>in</strong> terms of a mov<strong>in</strong>g frame we can def<strong>in</strong>e<br />

these 2-<strong>forms</strong> <strong>in</strong> S(CK n (ɛ)) for any ɛ.<br />

Remark 2.4.4. In [Par02] <strong>in</strong>variant 2-<strong>forms</strong> at CK n (ɛ) are def<strong>in</strong>ed <strong>in</strong> the same way as before<br />

(see Section 2.4.3).<br />

Proposition 2.4.5 ([Par02] Proposition 2.2.1). The algebra Ω ∗ (S(CK n (ɛ))) of R-valued <strong>in</strong>variant<br />

differential <strong>forms</strong> on the unit tangent bundle S(CK n (ɛ)) is generated by<br />

α, β, γ, θ 0 , θ 1 , θ 2 , θ s .<br />

From this proposition, Bernig and Fu def<strong>in</strong>e the families of (2n−1)-<strong>forms</strong> {β k,q } and {γ k,q }<br />

at S(C n ), but from the def<strong>in</strong>ition of θ 0 , θ 1 and θ 2 at S(CK n (ɛ)) these families of <strong>forms</strong> can be<br />

def<strong>in</strong>ed <strong>in</strong> the same way at S(CK n (ɛ)). By the previous proposition we have, as it is note <strong>in</strong><br />

[Par02], that all (2n − 1)-form <strong>in</strong>variant on S(CK n (ɛ)) such that they do not vanish over the<br />

normal bundle of a doma<strong>in</strong> Ω (cf. Lemma 1.3.12) are the ones given <strong>in</strong> the next def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition 2.4.6. Let k, q ∈ N be such that max{0, k − n} ≤ q ≤ k 2<br />

differential (2n − 1)-<strong>forms</strong> at S(CK n (ɛ)) are def<strong>in</strong>ed<br />

< n. Then, the follow<strong>in</strong>g<br />

where<br />

β k,q := c n,k,q β ∧ θ n−k+q<br />

0 ∧ θ k−2q−1<br />

1 ∧ θ q 2 , k ≠ 2q<br />

γ k,q := c n,k,q<br />

2 γ ∧ θn−k+q−1 0 ∧ θ k−2q<br />

1 ∧ θ q 2 , n ≠ k − q<br />

1<br />

c n,k,q :=<br />

q!(n − k + q)!(k − 2q)!ω 2n−k<br />

and ω 2n−k denotes the volume of the Euclidean ball of radius 1 and dimension 2n − k.<br />

Def<strong>in</strong>ition 2.4.7. Given a regular doma<strong>in</strong> Ω ⊂ CK n (ɛ), <strong>forms</strong> β k,q and γ k,q def<strong>in</strong>e the follow<strong>in</strong>g<br />

<strong>in</strong>variant valuations (see Section 2.4.3) <strong>in</strong> CK n (ɛ) (for max{0, k − n} ≤ q ≤ k 2 < n)<br />

∫<br />

∫<br />

B k,q (Ω) := β k,q (k ≠ 2q) and Γ k,q (Ω) = γ k,q (n ≠ k − q)<br />

N(Ω)<br />

N(Ω)<br />

where N(Ω) denotes the normal fiber bundle of Ω (see Def<strong>in</strong>ition 1.3.10).


2.4 Valuations on <strong>complex</strong> <strong>space</strong> <strong>forms</strong> 39<br />

In C n the previous valuations satisfy B k,q (Ω) = Γ k,q (Ω) s<strong>in</strong>ce dβ k,q = dγ k,q . For ɛ ≠ 0 no<br />

form β k,q has the same differential as γ k,q .<br />

The differential of the <strong>forms</strong> θ 0 , θ 1 and θ 2 is given <strong>in</strong> [BF08] for ɛ = 0. From the structure<br />

equations <strong>in</strong> CK n (ɛ) (cf. [KN69]), <strong>in</strong> the same way as <strong>in</strong> [Par02], we compute the differential<br />

for any ɛ.<br />

Lemma 2.4.8. In S(CK n (ɛ)) it is satisfied<br />

dα = −θ s , dθ 0 = −ɛ(α ∧ θ 1 + βθ s ),<br />

dβ = θ 1 , dθ 1 = dθ 2 = dθ s = 0<br />

dγ = 2θ 0 − 2ɛ(α ∧ β + θ 2 ).<br />

In the follow<strong>in</strong>g proposition we give the relation between valuations {B k,q (Ω)} and {Γ k,q (Ω)}<br />

<strong>in</strong> CK n (ɛ).<br />

Proposition 2.4.9. In CK n (ɛ), for any pair of <strong>in</strong>tegers (k, q) such that max{0, k − n} < q <<br />

k/2 < n it is satisfied<br />

c n,k,q<br />

Γ k,q (Ω) = B k,q (Ω) − ɛ B k+2,q+1 (Ω)<br />

c n,k+2,q+1<br />

(q + 1)(2n − k)<br />

= B k,q (Ω) − ɛ<br />

2π(n − k + q) B k+2,q+1(Ω).<br />

Proof. Denote by I the ideal generated by α, dα and all the exact <strong>forms</strong> <strong>in</strong> N(Ω) (see Def<strong>in</strong>ition<br />

1.3.10). If two <strong>forms</strong> λ and ρ of degree 2n − 1 co<strong>in</strong>cide modulo I, then by Lemma 1.3.12<br />

∫ ∫<br />

λ = ρ.<br />

Thus, it is enough to prove<br />

N(Ω)<br />

c n,k,q<br />

N(Ω)<br />

γ k,q ≡ β k,q − ɛ β k+2,q+1 mod I. (2.7)<br />

c n,k+2,q+1<br />

Consider the form η = (θ s − β ∧ γ) ∧ θ n−k+q−1<br />

0 θ k−2q−1<br />

1 θ q 2 . As dη is exact we have dη ≡ 0<br />

mod I. On the other hand, from the differentials given <strong>in</strong> Lemma 2.4.8 we obta<strong>in</strong><br />

dη ≡ −γθ n−k+q−1<br />

0 θ k−2q<br />

1 θ q 2 + 2βθn−k+q 0 θ k−2q−1<br />

1 θ q 2 − 2ɛβθn−k+q−1 0 θ k−2q−1<br />

1 θ q+1<br />

2 mod I.<br />

Us<strong>in</strong>g the def<strong>in</strong>ition of γ k,q and β k,q we get the relation (2.7).<br />

Remark 2.4.10. For n = 2, 3, the previous relations are given <strong>in</strong> [Par02].<br />

By the relation <strong>in</strong> Proposition 2.4.9, we def<strong>in</strong>e the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes <strong>in</strong> CK n (ɛ).<br />

Def<strong>in</strong>ition 2.4.11. For max{0, k−n} ≤ q ≤ k 2<br />

< n, we def<strong>in</strong>e the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes<br />

µ k,q <strong>in</strong> CK n (ɛ)<br />

{<br />

Bk,q (Ω) si k ≠ 2q<br />

µ k,q (Ω) :=<br />

(2.8)<br />

Γ 2q,q (Ω) si k = 2q.<br />

Remark 2.4.12. In C n , Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes form a basis of cont<strong>in</strong>uous valuations<br />

<strong>in</strong>variant under the isometry group of C n (cf. [BF08]).<br />

In the previous def<strong>in</strong>ition of µ k,q we take an arbitrary choice, despite of the relations <strong>in</strong><br />

Proposition 2.4.9. It would be <strong>in</strong>terest<strong>in</strong>g to know if there is a better choice, such that it<br />

satisfies some more natural geometric or algebraic properties.


40 Introduction to valuations<br />

2.4.3 Relation between Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes and the valuations given<br />

by Park<br />

We give, by completeness, the valuations def<strong>in</strong>ed by Park <strong>in</strong> CK n (ɛ).<br />

Def<strong>in</strong>ition 2.4.13. Denote θ 00 = −α ∧ β + θ 2 , θ 01 = −β ∧ γ + θ s , θ 10 = −α ∧ γ + θ 1 , θ 11 = θ 0 .<br />

Let κ = σ ∧ {a, b, c} where σ ∈ {α, β} and {a, b, c} = 1<br />

a!b!c! θa 11 ∧ θb 00 ∧ θc 10 . Then<br />

∫<br />

Φ κ (Ω) =<br />

N(Ω)<br />

are smooth valuation <strong>in</strong>variant under the action of the isometry group of CK n (ɛ).<br />

The relation between these valuations and the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes is given straightforward<br />

from the def<strong>in</strong>ition of each valuation.<br />

Proposition 2.4.14. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>. Then,<br />

1<br />

B k,q (Ω) =<br />

Φ β{n−k+q,q,k−2q−1} (Ω),<br />

(k − 2q)ω 2n−k<br />

1<br />

Γ k,q (Ω) =<br />

Φ γ{n−k+q−1,q,k−2q} (Ω).<br />

2(n − k + q)ω 2n−k<br />

κ<br />

2.4.4 Other curvature <strong>in</strong>tegrals<br />

Let M be a Kähler manifold of <strong>complex</strong> dimension n and suppose that S ⊂ M is a real<br />

hypersurface. Then, we can canonically def<strong>in</strong>e a distribution of <strong>complex</strong> dimension n − 1 <strong>in</strong><br />

the tangent fiber bundle of S <strong>in</strong> the follow<strong>in</strong>g way.<br />

Let N x be the normal fiber bundle of S at x. Let J be the <strong>complex</strong> structure <strong>in</strong> M. The<br />

vector JN x is a tangent vector to S at x. Consider the orthogonal vectors to JN x <strong>in</strong>side the<br />

tangent <strong>space</strong> of S at x. These form a <strong>complex</strong> sub<strong>space</strong> of dimension n − 1. Denote by D<br />

the distribution def<strong>in</strong>ed by these sub<strong>space</strong>s. Then, we def<strong>in</strong>e the mean curvature <strong>in</strong>tegrals<br />

restricted to the distribution D as<br />

Def<strong>in</strong>ition 2.4.15. Let S be a hypersurface of a Kähler manifold M of <strong>complex</strong> dimension n.<br />

If x ∈ S, we denote the second fundamental form of S at x by II x and the second fundamental<br />

form restricted to D by II x | D . The r-th mean curvature <strong>in</strong>tegral of S restricted at D, 1 ≤ r ≤<br />

2n − 2, is def<strong>in</strong>ed as<br />

( ) 2n − 2 −1 ∫<br />

Mr D (S) =<br />

σ r (II x | D )dx<br />

r<br />

where σ r (II x | D ) denotes the r-th symmetric elementary function of II x | D .<br />

Along this work, we use the idea of restrict<strong>in</strong>g mean curvature <strong>in</strong>tegrals to the distribution<br />

D. The first mean curvature <strong>in</strong>tegral restricted to D will play an important role, i.e. the<br />

<strong>in</strong>tegral of the trace of the second fundamental form restricted to the distribution. Also the<br />

<strong>in</strong>tegral over the normal curvature JN will have an important role. If Ω is a regular doma<strong>in</strong>,<br />

we have the follow<strong>in</strong>g relations<br />

∫<br />

(2n − 1)M 1 (∂Ω) − (2n − 2)M1 D (∂Ω) =<br />

∂Ω<br />

S<br />

k n (JN)dp = 2ω 2 Γ 2n−2,n−1 (Ω). (2.9)


2.4 Valuations on <strong>complex</strong> <strong>space</strong> <strong>forms</strong> 41<br />

2.4.5 Relation between the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes and the second<br />

fundamental form<br />

In this section we give another expression for the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes <strong>in</strong> terms of the<br />

second fundamental form.<br />

In the proof of Theorem 4.3.1 we shall use some properties, <strong>in</strong>terest<strong>in</strong>g by their own, of<br />

the expression of the <strong>in</strong>variant (2n − 1)-<strong>forms</strong> expressed <strong>in</strong> terms of the second fundamental<br />

form of ∂Ω, and not only <strong>in</strong> terms of the connection <strong>forms</strong> of a mov<strong>in</strong>g frame. In order to give<br />

this expression <strong>in</strong> terms of the second fundamental form (with respect to a fixed basis) it is<br />

necessary to consider the pull-back of the follow<strong>in</strong>g canonical map<br />

ϕ : ∂Ω −→ N(Ω)<br />

x ↦→ (x, N x ) . (2.10)<br />

Let us study some properties of ϕ ∗ (β k,q ) and ϕ ∗ (γ k,q ).<br />

Let x ∈ ∂Ω ⊂ CK n (ɛ) and let {e 1 = ϕ(x), e 1 = Je 1 , . . . , e n , e n = Je n } be a J-mov<strong>in</strong>g<br />

frame def<strong>in</strong>ed <strong>in</strong> a neighborhood of x. Consider the 1-<strong>forms</strong> {α i , β i , α 1j , β 1j }, and the 2-<strong>forms</strong><br />

{θ 0 , θ 1 , θ 2 , θ s } given at (2.6).<br />

Notation 2.4.16. In order to simplify the notation <strong>in</strong> the follow<strong>in</strong>g expressions we denote β i<br />

by α i<br />

and β 1i by α 1i<br />

and we def<strong>in</strong>e I := {1, 2, 2, . . . , n, n}.<br />

Now, us<strong>in</strong>g the relation between the connection <strong>forms</strong> α 1i , i ∈ I, and the second fundamental<br />

form<br />

α 1i = ∑ h ij α i , (2.11)<br />

j∈I<br />

we obta<strong>in</strong><br />

Lemma 2.4.17. In the previous notation,<br />

ϕ ∗ (β) = α 1 ,<br />

ϕ ∗ (γ) = ∑ h 1j α j ,<br />

j∈I<br />

n∑ ∑<br />

ϕ ∗ (θ 0 ) = h ij h il<br />

α j ∧ α l ,<br />

ϕ ∗ (θ 1 ) =<br />

ϕ ∗ (θ 2 ) =<br />

i=2 j,l∈I<br />

⎛<br />

n∑<br />

⎝ ∑ h ij<br />

α i ∧ α j − ∑<br />

j∈I<br />

l∈I<br />

i=2<br />

n∑<br />

α i ∧ α i<br />

.<br />

i=2<br />

h il α i<br />

∧ α l<br />

⎞<br />

⎠ ,<br />

On the other hand, each form ϕ ∗ (β k,q ) is a form of maximum degree, and, thus, a multiple<br />

of the volume element dx = α 1 ∧ α 2 ∧ α 2 ∧ · · · ∧ α n of ∂Ω. Thus, ϕ ∗ (β k,q ) is determ<strong>in</strong>ed by this<br />

multiple, which can be <strong>in</strong>terpreted as a polynomial with variables the entries of the second<br />

fundamental form h ij (with respect to the fixed J-mov<strong>in</strong>g frame).<br />

In [Par02], it is computed explicitly the pull-back of the <strong>forms</strong> β k,q and γ k,q for dimensions<br />

n = 2, 3. In the follow<strong>in</strong>g lemma we give some general properties for the pull-back of these<br />

<strong>forms</strong> for any dimension n.<br />

Lemma 2.4.18. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong> and let ϕ : ∂Ω → N(Ω) be the canonical<br />

map def<strong>in</strong>ed at (2.10). Let us fix a po<strong>in</strong>t x ∈ ∂Ω and a J-mov<strong>in</strong>g frame {e 1 = ϕ(x), e 1 =<br />

Je 1 , . . . , e n , e n = Je n } at x. If<br />

ϕ ∗ (β k,q ) = Q k,q dx,<br />

ϕ ∗ (γ k,q ) = P k,q dx


42 Introduction to valuations<br />

where dx is the volume element of ∂Ω, then<br />

1. Q k,q is a polynomial of degree 2n − k − 1 with variables the entries of the second fundamental<br />

form h ij = II(e i , e j ), i, j ∈ {2, 2, . . . , n};<br />

2. P k,q is a polynomial of degree 2n − k − 1 with variables the entries of the second fundamental<br />

form h ij = II(e i , e j ), i, j ∈ {1, 2, 2, . . . , n};<br />

3. each monomial of P k,q conta<strong>in</strong><strong>in</strong>g only entries of the form h ii also conta<strong>in</strong>s h 11 and<br />

exactly n + q − k − 1 factors of the form h jj h jj<br />

, i ∈ {1, 2, 2, . . . , n}, j ∈ {2, . . . , n};<br />

4. the polynomials P k,q and Q k,q can be written as a sum of m<strong>in</strong>ors of the second fundamental<br />

form with rank r = 2n − k − 1;<br />

5. among the m<strong>in</strong>ors described at 4. appear all m<strong>in</strong>ors centered at the diagonal with degree r<br />

conta<strong>in</strong><strong>in</strong>g h 11 for P k,q , and not conta<strong>in</strong><strong>in</strong>g h 11 for Q k,q . It also appears all non-centered<br />

m<strong>in</strong>ors such that the <strong>in</strong>dices of the rows and the <strong>in</strong>dices of the columns determ<strong>in</strong><strong>in</strong>g a<br />

m<strong>in</strong>or satisfy<br />

(a) conta<strong>in</strong> n − k + q <strong>in</strong>dices, <strong>in</strong> the case of Q k,q , and n − k + q − 1, <strong>in</strong> the case of P k,q ,<br />

such that, if the <strong>in</strong>dex j appears as an <strong>in</strong>dex <strong>in</strong> the rows (resp. columns), then the<br />

<strong>in</strong>dex j also appears as an <strong>in</strong>dex <strong>in</strong> the rows (resp. columns) of the m<strong>in</strong>or. We say<br />

that the <strong>in</strong>dex j is paired <strong>in</strong> the rows (or <strong>in</strong> the columns);<br />

(b) conta<strong>in</strong> k − 2q − 1 <strong>in</strong>dices non-paired neither <strong>in</strong> the rows nor <strong>in</strong> the columns for<br />

Q k,q , and k − 2q for P k,q ;<br />

(c) if the <strong>in</strong>dex j is <strong>in</strong> the non-paired <strong>in</strong>dices of the rows, then the <strong>in</strong>dex j is not <strong>in</strong> the<br />

<strong>in</strong>dex of the columns.<br />

Proof. From Lemma 2.4.17 we have<br />

and<br />

⎛<br />

⎞<br />

n∑ ∑<br />

ϕ ∗ (β k,q ) = c n,k,q α 1 ∧ ⎝ h ij h il<br />

α j ∧ α l<br />

⎠<br />

ϕ ∗ (γ k,q ) = c n,k,q<br />

2<br />

i=2 j,l∈I<br />

n+q−k<br />

⎛ ⎛<br />

⎞⎞<br />

n∑<br />

∧ ⎝ ⎝ ∑ h ij<br />

α i ∧ α j − ∑ h il α i<br />

∧ α l<br />

⎠⎠<br />

i=2 j∈I<br />

l∈I<br />

∧ (2.12)<br />

k−2q−1<br />

⎛ ⎞ ⎛<br />

⎞<br />

⎝ ∑ n+q−k−1<br />

n∑ ∑<br />

h 1j α j<br />

⎠ ∧ ⎝ h ij h il<br />

α j ∧ α l<br />

⎠ ∧<br />

j∈I<br />

i=2 j,l∈I<br />

⎛ ⎛<br />

⎞⎞<br />

n∑<br />

∧ ⎝ ⎝ ∑ h ij<br />

α i ∧ α j − ∑ h il α i<br />

∧ α l<br />

⎠⎠<br />

i=2 j∈I<br />

l∈I<br />

k−2q<br />

( n∑<br />

) q<br />

∧ α i ∧ α i<br />

,<br />

i=2<br />

( n∑<br />

) q<br />

∧ α i ∧ α i<br />

.<br />

i=2<br />

Thus, as ϕ ∗ (β k,q ) and ϕ ∗ (γ k,q ) are differential <strong>forms</strong> def<strong>in</strong>ed on ∂Ω of maximum degree, we<br />

have that ϕ ∗ (β k,q ) = Q k,q dx and ϕ ∗ (γ k,q ) = P k,q dx satisfy that Q k,p and P k,q are polynomials<br />

of degree 2n − k − 1. Moreover, polynomials P k,q cannot conta<strong>in</strong> h 11 s<strong>in</strong>ce this variable is<br />

multiplied by α 1 (see formula (2.11)), but this differential form is a common factor <strong>in</strong> the<br />

expression ϕ ∗ (β k,q ).


2.4 Valuations on <strong>complex</strong> <strong>space</strong> <strong>forms</strong> 43<br />

In order to prove 3. we observe that terms <strong>in</strong> the expression ϕ ∗ (γ k,q ) conta<strong>in</strong><strong>in</strong>g only entries<br />

of the type h ii are<br />

( n∑<br />

) n+q−k−1 ( n∑<br />

) k−2q ( n∑<br />

) q<br />

c n,k,q<br />

2 h 11 α 1 ∧ h ii h ii<br />

α i ∧ α i<br />

∧ h ii<br />

α i ∧ α i<br />

− h ii α i<br />

∧ α i ∧ α i ∧ α i<br />

.<br />

i=2<br />

i=2<br />

i=2<br />

So, the variable h 11 always appears and there are exactly n + q − k − 1 factors of the form<br />

h jj h jj<br />

, which come from ( ∑ n<br />

i=2 h iih ii<br />

α i ∧ α i<br />

) n+q−k−1 .<br />

A m<strong>in</strong>or of rank r of a matrix is def<strong>in</strong>ed choos<strong>in</strong>g r rows and r columns of the matrix and<br />

then tak<strong>in</strong>g the determ<strong>in</strong>ant of the square submatrix. To prove 4. and 5. we study <strong>in</strong> more<br />

detail the expression (2.12). (An analogous study can be done <strong>in</strong> the case ϕ ∗ (γ k,q ).) First, note<br />

that factors α 1 and ϕ ∗ (θ 2 ) do not give any term of the second fundamental form and, thus,<br />

they do not <strong>in</strong>fluence <strong>in</strong> the m<strong>in</strong>or (but they do <strong>in</strong>fluence <strong>in</strong> which m<strong>in</strong>ors can be constructed).<br />

Thus, we have to prove that the expression<br />

ϕ ∗ (θ n+q−k<br />

0 ) ∧ ϕ ∗ (θ k−2q−1<br />

1 ), (2.13)<br />

is a form of degree 2n−2q −2 where each term α i1 ∧α i2 ∧· · ·∧α i2n−2q−2 goes with a summation<br />

of m<strong>in</strong>ors.<br />

Develop<strong>in</strong>g (2.13) we have that it is equivalent to<br />

⎛<br />

⎞<br />

n∑<br />

n∑<br />

ϕ ∗ ⎝( α 1i ∧ α 1i<br />

) n+q−k ∧ ( (α j ∧ α 1j<br />

− α j ∧ α 1j )) k−2q−1 ⎠ .<br />

i=2<br />

j=2<br />

To develop this expression, first, we chose a := n + q − k values {i 1 , . . . , i a } for the <strong>in</strong>dex i<br />

of the first summation and b := k − 2q − 1 values {j 1 , . . . , j b } (with i k , j l ∈ {2, . . . , n}) for<br />

the <strong>in</strong>dex j of the second summation. Note that some <strong>in</strong>dexes can be repeated. Thus, we get<br />

( n−1<br />

n+q−k<br />

) (<br />

·<br />

n−1<br />

k−2q−1)<br />

summands<br />

ϕ ∗ (α 1i1 ∧ α 1i1<br />

∧ · · · ∧ α 1ia ∧ α 1ia<br />

∧ (α j1 ∧ α 1j1<br />

− α j1<br />

∧ α 1j1 ) ∧ · · · ∧ (α jb ∧ α 1jb<br />

− α jb<br />

∧ α 1jb )),<br />

which can be decomposed, for example, <strong>in</strong> the form<br />

ϕ ∗ (α 1i1 ∧ α 1i1<br />

∧ · · · ∧ α 1ia ∧ α 1ia<br />

∧ α j1 ∧ α 1j1<br />

∧ · · · ∧ α jb ∧ α 1jb<br />

)<br />

= α j1 ∧ · · · ∧ α jb ϕ ∗ (α 1i1 ∧ α 1i1<br />

∧ · · · ∧ α 1ia ∧ α 1ia<br />

∧ α 1j1<br />

∧ · · · ∧ α 1jb<br />

).<br />

From (2.11) the form <strong>in</strong> which we take pull-back can be expressed as the summation of the<br />

m<strong>in</strong>ors with rows given by the <strong>in</strong>dices I := {i 1 , i 1 , . . . , i a , i a , j 1 , . . . , j b }, and by columns each<br />

of the possible permutations of the elements without repetition among the <strong>in</strong>dexes <strong>in</strong> J :=<br />

I \ {1, j 1 , . . . , j b }. Note that <strong>in</strong>dex α 1 cannot be taken s<strong>in</strong>ce we are consider<strong>in</strong>g the form <strong>in</strong><br />

(2.12), which is multiplied by α 1 . (In the case of ϕ ∗ (γ k,q ) we do not have this restriction, and,<br />

so, can also appear m<strong>in</strong>ors with the term h 11 .)<br />

If we chose for J the same <strong>in</strong>dexes as <strong>in</strong> I, then we get all the m<strong>in</strong>ors centered at the<br />

diagonal. Condition 5.(c) is obta<strong>in</strong>ed directly from the fact that the <strong>in</strong>dices which determ<strong>in</strong>es<br />

the columns have to be <strong>in</strong> J , and if j k is an <strong>in</strong>dex <strong>in</strong> the rows, the <strong>in</strong>dex j k is not <strong>in</strong> J .<br />

Conditions 5.(a) and 5.(b) are obta<strong>in</strong>ed when we recall that we are not study<strong>in</strong>g the differential<br />

form <strong>in</strong> (2.13) but <strong>in</strong> (2.12), i.e. we have to take the product with α 1 ∧ ( ∑ n<br />

i=2 α i ∧ α i<br />

) q .<br />

But, if this product conta<strong>in</strong>s the form α k , then it also conta<strong>in</strong>s the form α k<br />

(except for k = 1).<br />

Thus, to complete the form <strong>in</strong> (2.13) to a (2n − 1)-differential form we have to take the products<br />

α k ∧ α k<br />

, so that, <strong>in</strong> order to not obta<strong>in</strong> a vanish<strong>in</strong>g term, the quantity of paired <strong>in</strong>dices<br />

<strong>in</strong> the rows and <strong>in</strong> the columns must be the same, and, thus, it also co<strong>in</strong>cides the quantity of<br />

no-paired <strong>in</strong>dices.


44 Introduction to valuations<br />

Remarks 2.4.19. 1. Each of the m<strong>in</strong>or goes with a constant depend<strong>in</strong>g on the number of<br />

permutations that allows us to obta<strong>in</strong> it. Thus, not all the m<strong>in</strong>ors have the same constant,<br />

but, for <strong>in</strong>stance all the m<strong>in</strong>ors (fixed β k,q or γ k,q ) centered at the diagonal have the same<br />

one.<br />

2. The degree of the polynomial given by β k,q or γ k,q does not depend on q, just on k, but<br />

two polynomials com<strong>in</strong>g from β k,q and β k,q ′ (or γ k,q and γ k,q ′) are dist<strong>in</strong>guished by the<br />

number of paired <strong>in</strong>dexes.<br />

Example 2.4.20. We give the explicit relation between some Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes and<br />

the second fundamental form.<br />

1. ϕ ∗ (γ 00 ) = c n,0,0<br />

2 ϕ∗ (γ ∧ θ0 n−1 ) = (n − 1)! c n,0,0<br />

det(II)dx = 1 det(II)dx.<br />

2<br />

2nω 2n<br />

2. ϕ ∗ (β 10 ) = c n,1,0 ϕ ∗ (β ∧ θ n−1<br />

0 ) = (n − 1)!c n,1,0 det(II| D )dx = 1<br />

ω 2n−1<br />

det(II| D )dx.<br />

3. ϕ ∗ (γ 2n−2,n−1 ) = c n,2n−2,n−1<br />

ϕ ∗ (γ∧θ n−1<br />

2<br />

4. ϕ ∗ (β 2n−2,n−2 ) = c n,2n−2,n−2 ϕ ∗ (β ∧ θ 1 ∧ θ n−2<br />

2 ) =<br />

2 ) = c n,2n−1,n−1(n − 1)!<br />

2<br />

k n (JN)dx = k n (JN) dx<br />

2ω 2<br />

.<br />

(n − 2)!<br />

(n − 2)!2ω 2<br />

tr(II| D )dx = tr(II| D ) dx<br />

2ω 2<br />

.<br />

5. ϕ ∗ (β 2n−1,n−1 ) = c n,2n−1,n−1 ϕ ∗ (β ∧ θ n−1<br />

2 ) = c n,2n−1,n−1 (n − 1)!α 1 ∧ α 1 ∧ · · · ∧ α n = dx 2 .


Chapter 3<br />

Average of the mean curvature<br />

<strong>in</strong>tegral<br />

For the real <strong>space</strong> <strong>forms</strong> (R n , S n and H n ), it is known that the reproductive property holds for<br />

mean curvature <strong>in</strong>tegrals. That is, given a regular doma<strong>in</strong> Ω, it is satisfied (cf. Example 2.2.3)<br />

∫<br />

M r<br />

(s)<br />

L s<br />

(∂Ω ∩ L s )dL s = cM r (∂Ω).<br />

On the other hand, by Section 2.3, this property may not hold <strong>in</strong> C n , when we <strong>in</strong>tegrate over<br />

the <strong>space</strong> of <strong>complex</strong> planes. Thus, it is natural to study, <strong>in</strong> C n , the value of<br />

∫<br />

L C s<br />

M (s)<br />

r (∂Ω ∩ L s )dL s .<br />

In the same way, we will study the value of this <strong>in</strong>tegral but <strong>in</strong> the other <strong>complex</strong> <strong>space</strong><br />

<strong>forms</strong>, CP n and CH n . Recall that we denote by CK n (ɛ) the <strong>space</strong> of constant holomorphic<br />

curvature 4ɛ.<br />

In this chapter we deduce the expression of the <strong>in</strong>tegral of M (s)<br />

1 (∂Ω ∩ L s) <strong>in</strong> terms of the<br />

mean curvature <strong>in</strong>tegral of the convex doma<strong>in</strong> M 1 (∂Ω) and the <strong>in</strong>tegral of the normal curvature<br />

<strong>in</strong> the direction JN, ∫ ∂Ω k n(JN) (see Theorem 3.3.2). We also f<strong>in</strong>d a partial result for the<br />

<strong>in</strong>tegral over any other mean curvature <strong>in</strong>tegral M r<br />

(s) (∂Ω∩L s ), 0 ≤ r ≤ 2s−1 (see Proposition<br />

3.2.2).<br />

3.1 Previous lemmas<br />

First of all, we state some lemmas that will be necessary <strong>in</strong> order to prove Theorem 3.3.2 and<br />

some other results.<br />

Lemma 3.1.1. Let V be a <strong>complex</strong> vector <strong>space</strong> of <strong>complex</strong> dimension 2 endowed with an <strong>in</strong>ner<br />

product 〈 , 〉 compatible with the <strong>complex</strong> structure J and let {e 1 , e 2 , e 3 , e 4 } be an orthonormal<br />

basis of V . Then 〈e a , Je b 〉 2 = 〈e c , Je d 〉 2 with {a, b, c, d} = {1, 2, 3, 4}.<br />

Proof. We express Je b and Je d <strong>in</strong> terms of the orthonormal, and we obta<strong>in</strong><br />

Je b = 〈Je b , e a 〉e a + 〈Je b , e c 〉e c + 〈Je b , e d 〉e d = Ae a + Be c + Ce d ,<br />

Je d = 〈Je d , e a 〉e a + 〈Je d , e b 〉e b + 〈Je d , e c 〉e c = De a + Ee b + F e c .<br />

45


46 Average of the mean curvature <strong>in</strong>tegral<br />

Now, us<strong>in</strong>g that 〈Je b , Je b 〉 = 〈Je d , Je d 〉 = 1, 〈Je b , Je d 〉 = 0 and 〈Je b , e d 〉 = −〈Je d , e b 〉, we<br />

get<br />

and<br />

A 2 + B 2 + C 2 = 1,<br />

D 2 + E 2 + F 2 = 1,<br />

AD + BF = 0,<br />

C = −E,<br />

A 2 + B 2 = D 2 + F 2 and A 2 D 2 = B 2 F 2 .<br />

F<strong>in</strong>ally, we substitute D 2 = A 2 + B 2 − F 2 <strong>in</strong> the second equation and we obta<strong>in</strong> A 2 = F 2 .<br />

Lemma 3.1.2. If u ∈ S 2n−3 , then<br />

∫<br />

S 2n−3 〈u, v〉dv = 0<br />

and<br />

∫<br />

S 2n−3 〈u, v〉 2 dv = ω 2n−2 .<br />

Proof. The first equality follows s<strong>in</strong>ce the <strong>in</strong>tegral is over an odd function. For the second one,<br />

we decompose v = cos θu + s<strong>in</strong> θw with w ∈ 〈u〉 ⊥ , then us<strong>in</strong>g polar coord<strong>in</strong>ates with respect<br />

to u, we have<br />

∫S 2n−3 〈u, v〉 2 dv = O 2n−4<br />

∫ π<br />

0<br />

cos 2 θ s<strong>in</strong> 2n−4 θdθ = O 2n−4<br />

O 2n−4+2+1<br />

O 2 O 2n−4<br />

= ω 2n−2 .<br />

where O n denotes the volume of the n-dimensional Euclidean sphere and ω n the volume of the<br />

the n-dimensional Euclidean ball.<br />

The follow<strong>in</strong>g lemma gives a generalized version of the Meus<strong>in</strong>er Theorem.<br />

Lemma 3.1.3. Let S ⊂ M be a hypersurface of class C 2 of a Riemannian manifold M, p ∈ S<br />

and L ⊂ T p M a vector sub<strong>space</strong>. We denote by II S the second fundamental form of S and<br />

by II L C the second fundamental form of C = S ∩ exp p L as a hypersurface of exp p L. We also<br />

denote u = T p S ∩ L. Then,<br />

σ i (II S | u ) = cos i θσ i (II| L C)<br />

where θ denotes the angle at p between a normal vector of S and a normal vector of C <strong>in</strong><br />

exp p L, and σ i (Q) denotes the i-th symmetric elementary function of the bil<strong>in</strong>eal form Q.<br />

Proof. If A ⊂ B ⊂ M are submanifolds, then we denote the second fundamental form of A as<br />

a submanifold of B by h B A : T pA × T p A → (T p A) ⊥ . If B = M, we just put h A <strong>in</strong>stead of h M A .<br />

Then, for all X, Y ∈ T p C<br />

h C (X, Y ) = h L C(X, Y ) + h L (X, Y ) = h L C(X, Y )<br />

s<strong>in</strong>ce the second fundamental form of L vanishes at p. Moreover,<br />

h C (X, Y ) = h S C(X, Y ) + h S (X, Y ).<br />

Let N be a normal vector to S. Note that h S C<br />

(X, Y ) is a multiple of a normal vector to C <strong>in</strong><br />

S, so 〈h S C (X, Y ), N〉 = 0 (for X, Y ∈ T pC).<br />

If X, Y ∈ T p C, then<br />

II S (X, Y ) := 〈h S (X, Y ), N〉 = 〈h C (X, Y ) − h S C(X, Y ), N〉<br />

= 〈h C (X, Y ), N〉 = 〈h L C(X, Y ), N〉 = 〈II L C(X, Y )n, N〉


3.1 Previous lemmas 47<br />

where n denotes a normal vector of C <strong>in</strong> L. So,<br />

II L C(X, Y ) = 1<br />

〈N, n〉 II S(X, Y ). (3.1)<br />

S<strong>in</strong>ce σ i (II L C ) is the sum of the m<strong>in</strong>ors of order i of IIL C , by replac<strong>in</strong>g by (3.1) each entry of the<br />

second fundamental form, we obta<strong>in</strong> the result.<br />

The follow<strong>in</strong>g lemma generalizes <strong>in</strong> C n a result given by Langev<strong>in</strong> and Shifr<strong>in</strong> [LS82] <strong>in</strong><br />

R n .<br />

Lemma 3.1.4. Let E be a <strong>complex</strong> vector <strong>space</strong> of <strong>complex</strong> dimension n and let II be a real<br />

bil<strong>in</strong>ear form def<strong>in</strong>ed on E. We denote by G C n,s the Grassmanian of s-dimensional <strong>complex</strong><br />

planes on E. Then,<br />

∫<br />

tr(II| V )dV = s vol(GC n,s)<br />

tr(II| E ).<br />

n<br />

G C n,s<br />

Proof. First, recall that<br />

is a fibration for each s ∈ {1, . . . , n − 1}.<br />

U(n − s) × U(s) −→ U(n) −→ G C n,s (3.2)<br />

We prove the case dim C V ≤ n 2<br />

by <strong>in</strong>duction on the <strong>complex</strong> dimension of V . The case<br />

dim C V > n 2<br />

can be proved us<strong>in</strong>g similar arguments.<br />

Suppose dim C V = 1, that is, s = 1. Then,<br />

∫<br />

G C n,1<br />

tr(II| V )dV =<br />

∫<br />

1<br />

tr(II|<br />

vol(U(n − 1))vol(U(1))<br />

V 1<br />

1<br />

)dU<br />

U(n)<br />

s<strong>in</strong>ce tr(II| V 1<br />

1<br />

) is constant along the fiber. We denote by V1 1 the <strong>complex</strong> vector sub<strong>space</strong><br />

generated by the first column of the matrix U ∈ U(n). In general, for U ∈ U(n), we will<br />

denote by Va b the <strong>complex</strong> vector sub<strong>space</strong> generated by the columns b to b + a − 1. The<br />

subscript a denotes the dimension of Va b , or equivalently, the number of columns we consider<br />

and the upperscript b denotes from which column we start to consider them. Then<br />

∫<br />

U(n)<br />

tr(II| V 1<br />

1<br />

)dU = 1 n<br />

= 1 n<br />

∫<br />

∫<br />

U(n)<br />

U(n)<br />

(tr(II| V 1<br />

1<br />

) + tr(II| V 2<br />

1<br />

) + · · · + tr(II| V n<br />

1<br />

))dU<br />

tr(II| E )dU = vol(U(n)) tr(II| E ).<br />

n<br />

Thus,<br />

∫<br />

G C n,1<br />

tr(II| V )dV =<br />

vol(U(n))<br />

nvol(U(n − 1))vol(U(1)) tr(II| E) = vol(GC n,1 ) tr(II| E ).<br />

n<br />

Suppose now that the result is true till dim C V = r −1. We shall prove it for dim C V = r ≤<br />

n<br />

2 . If R denotes the rema<strong>in</strong>der of n r<br />

, then R < r and we can apply the <strong>in</strong>duction hypothesis <strong>in</strong>


48 Average of the mean curvature <strong>in</strong>tegral<br />

R at equality (*). Thus, us<strong>in</strong>g similar arguments as before, we obta<strong>in</strong><br />

∫<br />

G C n,r<br />

tr(II| V ) =<br />

∫<br />

1<br />

tr(II|<br />

vol(U(n − r))vol(U(r))<br />

V 1 r<br />

)<br />

U(n)<br />

1<br />

(tr(II| V 1 r<br />

) + tr(II| V 2 r<br />

) + · · · + tr(II| V<br />

n−R−r+1<br />

=<br />

vol(U(n − r))vol(U(r))⌊ n r<br />

∫U(n) ⌋<br />

( ∫<br />

1<br />

=<br />

vol(U(n − r))vol(U(r))⌊ n r ⌋ U(n)<br />

(∗)<br />

=<br />

∫<br />

tr(II| E ) −<br />

U(n)<br />

tr(II| V<br />

n−R+1)<br />

R<br />

vol(U(n))tr(II| E ) − vol(U(n − R))vol(U(R)) ∫ G<br />

tr(II| C VR )<br />

n,R<br />

=<br />

vol(U(n − r))vol(U(r))⌊ n r<br />

(<br />

⌋ )<br />

vol(U(n)) − vol(U(n − R))vol(U(R))vol(G C n,R ) R n<br />

tr(II| E )<br />

= vol(G C r<br />

n,r)<br />

n − R<br />

= vol(G C n,r) r n tr(II| E)<br />

and the result follows when 2s ≤ n.<br />

vol(U(n − r))vol(U(r))⌊ n r<br />

(<br />

⌋<br />

1 − R )<br />

tr(II| E )<br />

n<br />

)<br />

r<br />

))<br />

3.2 <strong>Integral</strong> of the r-th mean curvature <strong>in</strong>tegral over the <strong>space</strong><br />

of <strong>complex</strong> s-planes<br />

Along this chapter we follow some conventions which we state <strong>in</strong> the follow<strong>in</strong>g paragraphs.<br />

We denote by S ⊂ CK n (ɛ) a hypersurface of class C 2 , compact and oriented (possibly with<br />

boundary). Given a <strong>complex</strong> s-plane L s <strong>in</strong>tersect<strong>in</strong>g S, it is said that L s is <strong>in</strong> generic position<br />

if S ∩ L s is a submanifold of dimension 2s − 1 <strong>in</strong> CK n (ɛ). For hypersurfaces of class C 2 , the<br />

subset of generic planes (<strong>in</strong>tersect<strong>in</strong>g S) has full measure. Thus, we suppose that each <strong>complex</strong><br />

s-plane is <strong>in</strong> generic position. Note that S ∩ L s (if L s is a <strong>complex</strong> s-plane <strong>in</strong> generic position)<br />

is a hypersurface <strong>in</strong> L s<br />

∼ = CK s (ɛ).<br />

Suppose that N denotes a unit normal vector field on S. In this chapter we take, <strong>in</strong> S ∩ L s<br />

as a submanifold <strong>in</strong> S, the normal vector field Ñ such that the angle between N and Ñ is<br />

acute. Along the proofs <strong>in</strong> this chapter, we denote e s := ±JÑ.<br />

Note that if p ∈ S ∩ L s and Ñp is the chosen normal vector field <strong>in</strong> S ∩ L s <strong>in</strong>side L s then<br />

JÑ ∈ T p(S ∩ L s ). Indeed, L s<br />

∼ = CK s (ɛ), thus, the same structure for hypersurfaces hols <strong>in</strong>side<br />

L s .<br />

We denote by E ⊂ T p CK n (ɛ), p ∈ S ∩ L s , the orthogonal sub<strong>space</strong> to the <strong>space</strong> generated<br />

by {N, JN, Ñ, JÑ}. Note that exp p(E) ∼ = CK n−2 (ɛ) and it is univocally determ<strong>in</strong>ed for each<br />

L s .<br />

The fact stated <strong>in</strong> the follow<strong>in</strong>g remark is used implicitly along this chapter, specially to<br />

def<strong>in</strong>e the mov<strong>in</strong>g frames g and g ′ <strong>in</strong> the proof of the next proposition.<br />

Remark 3.2.1. Let V n be an n-dimensional Hermitian <strong>space</strong> with <strong>complex</strong> structure J, H ⊂ V n<br />

a real hyperplane and W s ⊂ V n a <strong>complex</strong> sub<strong>space</strong> of dimension s.<br />

Consider the sub<strong>space</strong> H ∩ W s and denote by N ′ an orthonormal vector to H ∩ W s <strong>in</strong> W s ,<br />

and D ′ = 〈N ′ , JN ′ 〉 ⊥ ∩ W s .<br />

Consider also the sub<strong>space</strong>s H ∩ Ws<br />

⊥<br />

<strong>in</strong> Ws ⊥ , and D ′′ = 〈N ′′ , JN ′′ 〉 ⊥ ∩ W s ′′ .<br />

and denote by N ′′ an orthonormal vector to H ∩ W ⊥ s


3.2 <strong>Integral</strong> of the r-th mean curvature <strong>in</strong>tegral 49<br />

Denote by N an orthonormal vector to H <strong>in</strong> V , and D = 〈N, JN〉 ⊥ . Then,<br />

V n = D⊥〈JN〉 R ⊥〈N〉 R<br />

= D ′ ⊥〈JN ′ 〉 R ⊥〈N ′ 〉 R ⊥D ′′ ⊥〈JN ′′ 〉 R ⊥〈N ′′ 〉 R .<br />

Indeed, from section 2.4.4 given a real hyperplane W <strong>in</strong> a Hermitian <strong>space</strong> V there exists<br />

a canonical decomposition of V as<br />

V = D⊥〈JN〉 R ⊥〈N〉 R ,<br />

where N is an orthogonal vector to W <strong>in</strong> V . Apply<strong>in</strong>g this fact to V = W s and to V = W ⊥ s<br />

we get the result.<br />

The follow<strong>in</strong>g proposition shall be essential to prove Theorem 3.3.2 and other results,<br />

s<strong>in</strong>ce it gives a first expression of the <strong>in</strong>tegral over the <strong>space</strong> of <strong>complex</strong> s-planes of the mean<br />

curvature <strong>in</strong>tegral <strong>in</strong> terms of an <strong>in</strong>tegral on the boundary of the doma<strong>in</strong>.<br />

Proposition 3.2.2. Let S ⊂ CK n (ɛ) be a compact (possibly with boundary) hypersurface of<br />

class C 2 oriented by a normal vector N, and let r, s ∈ N such that 1 ≤ s ≤ n and 0 ≤ r ≤ 2s−1.<br />

Then<br />

∫<br />

( ) 2s − 1 −1 ∫ ∫ ( ∫ )<br />

M r (s)<br />

|〈JN, e s 〉|<br />

(S∩L s )dL s =<br />

L C r<br />

s<br />

S( 2s−r<br />

RP 2n−2 G C (1 − 〈JN, e<br />

n−2,s−1<br />

s 〉 2 ) s−1 σ r(p; e s ⊕ V )dV<br />

)de s dp,<br />

where e s ∈ T p S unit vector, V denotes a <strong>complex</strong> (s−1)-plane by p conta<strong>in</strong>ed <strong>in</strong> {N, JN, e s , Je s } ⊥ ,<br />

σ r (p; e s ⊕ V ) denotes the r-th symmetric elementary function of the second fundamental form<br />

of S restricted to the real sub<strong>space</strong> e s ⊕V and the <strong>in</strong>tegration over RP 2n−2 denotes the projective<br />

<strong>space</strong> of the unit tangent <strong>space</strong> of the hypersurface.<br />

Remark 3.2.3. Us<strong>in</strong>g the previous remarks, it follows that the product |〈JN, e s 〉| <strong>in</strong> the last<br />

proposition gives the cos<strong>in</strong>e of the acute angle between the normal vector to the hypersurface<br />

S <strong>in</strong> CK n (ɛ) and a normal vector to S ∩ L s <strong>in</strong> L s , that is,<br />

|〈JN, e s 〉| = |〈N, Ñ〉|.<br />

Proof. Let L s be a <strong>complex</strong> s-plane such that S ∩ L s ≠ ∅ and let p ∈ S ∩ L s . We denote<br />

by ˜σ r the r-th symmetric elementary function of the second fundamental form of S ∩ L s as a<br />

hypersurface of L s . Then, by def<strong>in</strong>ition<br />

∫<br />

( ) 2s − 1 −1 ∫ ∫<br />

M r (s) (S ∩ L s )dL s =<br />

˜σ r (s)dx dL s .<br />

r S∩L s≠∅ S∩L s<br />

L C s<br />

We shall prove the result us<strong>in</strong>g mov<strong>in</strong>g frames adapted to S ∩ L s , L s or S.<br />

Let g = {e 1 , e 1 = Je 1 , e 2 , e 2 = Je 2 , . . . , e s , w s , e s+1 , e s+1 = Je s+1 . . . , e n , N} be a mov<strong>in</strong>g<br />

frame adapted to S ∩ L s and S (cf. Remark 3.2.1). That is, {e 1 , e 1 , . . . , e s } is an orthonormal<br />

basis of T p (S ∩L s ), {e s+1 , e s+1 , . . . , e n } is an orthonormal basis of T p S ∩(T p L s ) ⊥ , N is a normal<br />

vector field to T S and w s completes to an orthonormal basis of T p CK n (ɛ). We denote by<br />

{ω 1 , ω 1 , . . . , ω s−1 , ω s−1 , ω s , ω˜s , ω s+1 , ω s+1 , . . . , ω n , ωñ}<br />

the dual basis of the vectors <strong>in</strong> g and by {ω ij }, i, j ∈ {1, 1, . . . , s, ˜s, s + 1, s + 1, . . . , n, ñ}, the<br />

connection <strong>forms</strong> (cf. (1.15)).<br />

Let g ′ = {e ′ 1 = e 1, e ′ 1 = Je 1, e ′ 2 = e 2, e ′ 2 = Je 2, ..., e ′ s = e s , e ′ s = Je s, e ′ s+1 = e s+1, e ′ s+1 =<br />

Je s+1 , ..., e ′ n = e n , e ′ n = Je n} be a mov<strong>in</strong>g frame adapted to S ∩ L s and L s . That is,


50 Average of the mean curvature <strong>in</strong>tegral<br />

{e ′ 1 , e′ 1 , ..., e′ s} is an orthonormal basis of T p (L s ∩ S) and {e ′ 1 , e′ 1 , . . . , e′ s, e s ′ } is an orthonormal<br />

basis of T p L s . Denote by<br />

{ω 1, ′ ω ′ 1 , . . . , ω′ n, ω n}<br />

′<br />

the dual basis of vectors <strong>in</strong> g ′ and by {ω ij ′ } the connection <strong>forms</strong>, i, j ∈ {1, 1, . . . , n, n}.<br />

As the base g and g ′ are constituted by orthonormal vectors, we can easily give the relation<br />

among the elements <strong>in</strong> the frame g ′ and the ones <strong>in</strong> g just express<strong>in</strong>g the vectors <strong>in</strong> g ′ <strong>in</strong><br />

coord<strong>in</strong>ates with respect to the vectors <strong>in</strong> g:<br />

e ′ s = Je s = 〈Je s , w s 〉w s + 〈Je s , N〉N,<br />

e ′ n = Je n = 〈Je n , w s 〉w s + 〈Je n , N〉N<br />

and e ′ j = e j when j ∈ {1, 1, . . . , s − 1, s − 1, s, s + 1, s + 1, . . . , n − 1, n − 1, n}.<br />

Then { ω<br />

′<br />

j = ω j , if j ≠ s, n,<br />

and<br />

⎧<br />

⎨<br />

⎩<br />

ω ′ n<br />

= 〈Je n , w s 〉ω˜s + 〈Je n , N〉ωñ<br />

ω is ′ = 〈Je s , w s 〉ω i˜s + 〈Je s , N〉ω iñ , if i ≠ s, n<br />

ω <strong>in</strong> ′ = 〈Je n , w s 〉ω i˜s + 〈Je n , N〉ω iñ , if i ≠ s, n<br />

ω ij ′ = ω ij , if i, j ≠ s, n.<br />

(3.3)<br />

(3.4)<br />

From now on, <strong>in</strong> order to simplify the notation, we omit the absolute value <strong>in</strong> densities.<br />

The expression of dx (the density of S ∩ L s ), dL s and dL s[p] <strong>in</strong> terms of ω ′ is<br />

dx = ω ′ 1 ∧ ω ′ 1 ∧ · · · ∧ ω′ s,<br />

dL s = ω ′ s+1 ∧ ω ′ s+1 ∧ · · · ∧ ω′ n ∧ ω ′ n ∧<br />

dL s[p] =<br />

∧<br />

ω ij<br />

′<br />

i=1,2,...,s<br />

j=s+1,s+1,...,n,n<br />

∧<br />

ω ′ ij,<br />

i=1,2,...,s<br />

j=s+1,s+1,...,n,n<br />

and the expression of dp (the density of S) <strong>in</strong> terms of ω is dp = ω 1 ∧ ω 1 ∧ · · · ∧ ω n .<br />

On the other hand, by Lemma 3.1.1 it is satisfied<br />

|〈Je n , w s 〉| = |〈Je s , N〉|. (3.5)<br />

Indeed, vectors {e s , w s , e n , N} are an orthonormal basis of a 2-dimensional <strong>complex</strong> plane, the<br />

orthogonal complement of the <strong>space</strong> generated by {e 1 , Je 1 , . . . , e s−1 , Je s−1 , e s+1 , Je s+1 , . . . , e n−1 , Je n−1 }.<br />

By relations (3.3) and (3.5) we get<br />

s<strong>in</strong>ce ωñ vanishes on T S.<br />

Then, by Lemma 3.1.3,<br />

∫<br />

M (s)<br />

S∩L s≠∅<br />

dx ∧ dL s = |〈Je n , w s 〉|dL s[p] ∧ dp = |〈JN, e s 〉|dL s[p] ∧ dp (3.6)<br />

( ) 2s − 1 −1 ∫<br />

r (S ∩ L s )dL s =<br />

r<br />

( ) 2s − 1 −1 ∫<br />

=<br />

r<br />

S<br />

S<br />

∫<br />

|〈JN, e s 〉|˜σ r (p)dL s[p] dp<br />

L s[p]<br />

∫<br />

|〈JN, e s 〉|<br />

L s[p]<br />

|〈N, Je s 〉| r σ r(p)dL s[p] dp.<br />

Note that <strong>in</strong> the last <strong>in</strong>tegrand we consider the absolut value <strong>in</strong> the denom<strong>in</strong>ator to be<br />

sure that we consider the acute angle between the two <strong>in</strong>tersect<strong>in</strong>g sub<strong>space</strong>. This is not a


3.2 <strong>Integral</strong> of the r-th mean curvature <strong>in</strong>tegral 51<br />

priori true s<strong>in</strong>ce e s is any vector and, thus, not always the vector Je s (a normal vector to<br />

S ∩ L s ⊂ L s ) has the desired orientation.<br />

Now, we shall express dL s[p] <strong>in</strong> terms of dV ∧ de s where dV denotes the volume element of<br />

the Grassmannian G C n−2,s−1 and de s the volume element of S 2n−2 . Fixed p, def<strong>in</strong>e the manifold<br />

M = {(e s , V ) | e s ∈ T p S unit , V ∈ L s conta<strong>in</strong><strong>in</strong>g p and orthogonal to {N, JN, e s , Je s }},<br />

which locally co<strong>in</strong>cides with S 2n−2 × G C n,s−1 . Consider the fiber bundle<br />

φ : M −→ L C s[p]<br />

(e s , V ) ↦→ exp p {e s , Je s , V }<br />

with fiber G C n,s. This is a double cover<strong>in</strong>g of L C s[p]<br />

s<strong>in</strong>ce vectors v and −v give the same <strong>complex</strong><br />

s-plane.<br />

The pull-back of dL s[p] by the last mapp<strong>in</strong>g give the desired relation among the densities.<br />

The expression of de s <strong>in</strong> terms of ω and the expression of dV <strong>in</strong> terms of ω ′ are<br />

∧<br />

de s =<br />

ω sj ,<br />

By (3.4) and (3.7) we have<br />

∧<br />

dL s[p] =<br />

dV =<br />

ω ij<br />

′<br />

i=1,2,...,s<br />

j=s+1,s+1,...,n,n<br />

= dV ∧<br />

∧<br />

i=1,2,...,s−1<br />

j=1,1,...,s−1,s−1,˜s,s+1,s+1,...,n<br />

∧<br />

i=1,2,...,s−1<br />

j=s+1,s+1,...,n−1,n−1<br />

ω <strong>in</strong> ∧<br />

∧<br />

i=1,2,...,s<br />

ω ′ ij. (3.7)<br />

ω ′ <strong>in</strong> ∧<br />

Next, we relate ∧ ω <strong>in</strong> ′ ∧ ω<br />

′<br />

<strong>in</strong> with ∧ ω sj . From (3.4) follows<br />

∧<br />

∧<br />

ω <strong>in</strong> ′ = |〈Je n , w s 〉| s<br />

i=1,2,...,s<br />

i=1,2,...,s<br />

∧<br />

j=s+1,s+1,...,n−1,n<br />

and also us<strong>in</strong>g that ω <strong>in</strong> ′ = ω′ <strong>in</strong> s<strong>in</strong>ce ω′ <strong>in</strong> = 〈de′ i , e′ n〉 = 〈dJe ′ i , Je′ n〉 = ω ′ we obta<strong>in</strong><br />

i,n<br />

∧<br />

∧<br />

ω <strong>in</strong> ′ = ω ′ <strong>in</strong> =<br />

∧<br />

〈Je n , w s 〉ω i˜s<br />

In order to study<br />

i=1,2,...,s−1<br />

i=1,2,...,s−1<br />

= |〈Je n , w s 〉| s−1 ∧<br />

∧<br />

i=1,1,...,s−1,s−1<br />

i=1,2,...,s−1<br />

i=1,2,...,s−1<br />

ω i˜s ,<br />

we use e s = Je s = 〈Je s , w s 〉w s + 〈Je s , N〉N and we obta<strong>in</strong><br />

∧<br />

∧<br />

ω is =<br />

ω is<br />

=<br />

i=1,1,...,s−1,s−1<br />

i=1,1,...,s−1,s−1<br />

= 〈Je s , w s 〉 2(s−1) ∧<br />

ω i˜s<br />

.<br />

ω i˜s<br />

∧<br />

i=1,1,...,s−1,s−1<br />

i=1,1,...,s−1,s−1<br />

ω i˜s .<br />

ω is<br />

ω sj .


52 Average of the mean curvature <strong>in</strong>tegral<br />

Thus,<br />

and<br />

∧<br />

i=1,1,...,s−1,s−1<br />

ω i˜s = 〈Je s , w s 〉 −2(s−1)<br />

∧<br />

i=1,1,...,s−1,s−1<br />

dL s[p] = |〈Je n, w s 〉| 2s−1<br />

〈Je s , w s 〉 2(s−1) dV ∧ de s. (3.8)<br />

Us<strong>in</strong>g |〈Je n , w s 〉| = |〈JN, e s 〉| and 〈Je s , w s 〉 2 = 1 − 〈JN, e s 〉 2 we get the result.<br />

3.3 Mean curvature <strong>in</strong>tegral<br />

First we give an expression for the <strong>in</strong>tegral over the <strong>space</strong> of <strong>complex</strong> r-planes of the <strong>in</strong>tegral<br />

of the normal curvature <strong>in</strong> the direction JN (see (2.9)). By Example 2.4.20 we have that this<br />

<strong>in</strong>tegral is a smooth valuation <strong>in</strong> CK n (ɛ). Moreover, it is not a multiple of the mean curvature<br />

<strong>in</strong>tegral.<br />

Theorem 3.3.1. Let S ⊂ CK n (ɛ) be compact oriented (possibly with boundary) hypersurface<br />

of class C 2 oriented by a normal vector N, and s ∈ {1, . . . , n − 1}. Then<br />

∫<br />

)<br />

˜kn<br />

(∫S∩L (JÑ)dx dL s<br />

s<br />

L C s<br />

= vol(G C n−2,s−1) ω ( )<br />

2n−2 n −1 ( ∫<br />

)<br />

2sn − s − n<br />

k n (JN) + (2n − 1)M 1 (S) ,<br />

2s s n − s S<br />

where ˜k n (JÑ) the normal curvature of S ∩ L s <strong>in</strong> the direction JÑ, k n(JN) the normal curvature<br />

of JN <strong>in</strong> CK n (ɛ) and ω 2n−2 denotes the volume of the unit ball <strong>in</strong> the standard Euclidean<br />

<strong>space</strong> of dimension 2n − 2.<br />

Proof. Denote JÑ by e s.<br />

By Lemma 3.1.3 we have ˜k n (JÑ) = k n(e s )<br />

. Us<strong>in</strong>g equalities (3.6) and (3.8) we obta<strong>in</strong><br />

〈JN, e s 〉<br />

I =<br />

∫<br />

=<br />

∫L s<br />

∫<br />

S<br />

S∩L s<br />

∫ ∫RP 2n−2<br />

˜ kn (JÑ)dxdL s<br />

G C n−2,s−1<br />

ω is<br />

〈JN, e s 〉 2s<br />

(1 − 〈JN, e s 〉 2 ) s−1 k n (e s )<br />

〈JN, e s 〉 dV de sdp<br />

As the <strong>in</strong>tegral over G C n−2,s−1 is <strong>in</strong>dependent of V , it follows<br />

∫<br />

I = vol(G C 〈JN, e s 〉<br />

n−2,s−1)<br />

S<br />

∫RP 2s−1<br />

2n−2 (1 − 〈JN, e s 〉 2 ) s−1 k n(e s )de s dp. (3.9)<br />

In order to compute the <strong>in</strong>tegral over RP 2n−2 , we use polar coord<strong>in</strong>ates and express the<br />

normal curvature of e s <strong>in</strong> terms of the pr<strong>in</strong>cipal curvatures of T p S.<br />

That is, if {f 1 , . . . , f 2n−1 } is an orthonormal basis of pr<strong>in</strong>cipal directions of T p S then<br />

e s = ∑ 2n−1<br />

j=1 〈e s, f j 〉f j , and<br />

2n−1<br />

∑<br />

2n−1<br />

∑<br />

k n (e s ) = 〈e s , f j 〉 2 k n (f j ) = 〈e s , f j 〉 2 k j .<br />

j=1<br />

On the other hand, we consider a polar coord<strong>in</strong>ates system θ 1 , θ 2 with respect to JN<br />

def<strong>in</strong>ed by<br />

|〈JN, e s 〉| = cos θ 1 , (3.10)<br />

j=1


3.3 Mean curvature <strong>in</strong>tegral 53<br />

and us<strong>in</strong>g spherical trigonometry,<br />

〈e s , f j 〉 = cos θ 1 cos(JN, f j ) + s<strong>in</strong> θ 1 s<strong>in</strong>(JN, f j ) cos(e s , JN, f j )<br />

= cos θ 1 cos α j + s<strong>in</strong> θ 1 s<strong>in</strong> α j cos θ 2 (3.11)<br />

where cos(e s , JN, f j ) denotes the cos<strong>in</strong>e of the spherical angle with vertex JN. Note that α j<br />

are constants when the po<strong>in</strong>t is fixed.<br />

Then, from the relations<br />

Γ(s)Γ(n − s)<br />

Γ(n + 1)<br />

=<br />

)<br />

1 n −1<br />

and<br />

s(n − s)(<br />

s<br />

Γ(s)Γ(n − s + 1)<br />

Γ(n + 1)<br />

= 1 s<br />

( n<br />

s<br />

) −1<br />

,<br />

we get<br />

∫RP 2n−2 |〈JN, e s 〉| 2s−1<br />

=<br />

+<br />

=<br />

2n−1<br />

∑<br />

j=1<br />

∫ π ∫S 2n−4<br />

2n−1<br />

∑<br />

j=1<br />

= ω 2n−2<br />

2s<br />

(1 − 〈JN, e s 〉 2 ) s−1 k n(e s )de s<br />

k j<br />

( ∫ S 2n−3 ∫ π/2<br />

0<br />

0<br />

cos 2s−1 θ 1<br />

s<strong>in</strong> 2s−2 θ 1<br />

cos 2 θ 1 cos 2 α j s<strong>in</strong> 2n−3 θ 1 dθ 1 dS 2n−3 +<br />

cos 2 θ 2 s<strong>in</strong> 2n−4 θ 2<br />

∫ π/2<br />

(<br />

k j O 2n−3 cos 2 (2sn − s − n)Γ(s)Γ(n − s)<br />

α j +<br />

4(n − 1)Γ(n + 1)<br />

( n −1 2n−1 ∑<br />

( )<br />

2sn − n − s<br />

k j cos<br />

s) 2 α j + 1 .<br />

n − s<br />

j=1<br />

Integrat<strong>in</strong>g over S and us<strong>in</strong>g<br />

we obta<strong>in</strong> the stated result.<br />

0<br />

k n (JN) =<br />

2n−1<br />

∑<br />

j=1<br />

cos 2s−1 θ<br />

)<br />

1<br />

s<strong>in</strong> 2s−2 s<strong>in</strong> 2 θ 1 s<strong>in</strong> 2 α j s<strong>in</strong> 2n−3 θ 1 dθ 1 dθ 2 dS 2n−4 + 0<br />

θ 1<br />

)<br />

k j 〈JN, f j 〉 2 =<br />

Γ(s)Γ(n − s + 1)<br />

4(n − 1)Γ(n + 1)<br />

2n−1<br />

∑<br />

j=1<br />

k j cos 2 α j<br />

Theorem 3.3.2. Let S ⊂ CK n (ɛ) be a compact (possibly with boundary) hypersurface of class<br />

C 2 oriented by a normal vector N, and let s ∈ {1, ..., n − 1}. Then<br />

∫<br />

L C s<br />

M (s)<br />

1 (S∩L s)dL s = ω 2n−2vol(G C n−2,s−1 ) ( n −1 (<br />

(2n − 1)<br />

2s(2s − 1) s) 2ns − n − s ∫<br />

M 1 (S) +<br />

n − s<br />

where k n (JN) denotes the normal curvature <strong>in</strong> the direction JN ∈ T S.<br />

L C s<br />

S<br />

)<br />

k n (JN)<br />

Proof. By Proposition 3.2.2 and Lemma 3.1.4 we have<br />

∫<br />

M (s)<br />

1 (S ∩ L s)dL s = 1 ∫ ∫<br />

|〈JN, e s 〉|<br />

2s − 1<br />

∫RP 2s−1<br />

2n−2 (1 − 〈JN, e s 〉 2 ) s−1 σ 1(p; e s , V )dV de s dp<br />

= 1<br />

2s − 1<br />

∫<br />

S<br />

= vol(GC n−2,s−1 )<br />

2s − 1<br />

S<br />

S<br />

G C n−2,s−1<br />

|〈JN, e s 〉|<br />

∫RP 2s−1 ∫<br />

2n−2 (1 − 〈JN, e s 〉 2 ) s−1 (tr(II| V ) + II(e s , e s ))dV de s dp<br />

G C n−2,s−1<br />

∫<br />

|〈JN, e s 〉|<br />

∫RP 2s−1 ( )<br />

s − 1<br />

2n−2 (1 − 〈JN, e s 〉 2 ) s−1 n − 2 tr(II| E) + k n (e s ) de s dp,<br />

where E = 〈N, JN, e s , Je s 〉 ⊥ .


54 Average of the mean curvature <strong>in</strong>tegral<br />

Note that if s = 1 then dim V = 0. Although the <strong>in</strong>tegral ∫ G<br />

tr(II|V )dV has no<br />

C<br />

n−2,s−1<br />

sense, last equality above rema<strong>in</strong>s true s<strong>in</strong>ce s−1<br />

n−2 trII| E = 0.<br />

We shall study the follow<strong>in</strong>g <strong>in</strong>tegrals<br />

J E = vol(GC n−2,s−1 )<br />

2s − 1<br />

J s = vol(GC n−2,s−1 )<br />

2s − 1<br />

∫<br />

s − 1<br />

n − 2<br />

∫<br />

S<br />

S<br />

∫RP 2n−2 |〈JN, e s 〉| 2s−1<br />

(1 − 〈JN, e s 〉 2 ) s−1 tr(II| E)de s dp<br />

∫RP 2n−2 |〈JN, e s 〉| 2s−1<br />

(1 − 〈JN, e s 〉 2 ) s−1 k n(e s )de s dp.<br />

The second <strong>in</strong>tegral is the same (except for a constant factor) as the <strong>in</strong>tegral (3.9) <strong>in</strong> Proposition<br />

3.3.1. Thus, we know<br />

J s = O 2n−3vol(G C n−2,s−1 ) ( ) n −1 ( ∫<br />

)<br />

2sn − s − n<br />

k n (JN) + (2n − 1)M 1 (S) .<br />

4s(2s − 1)(n − 1) s n − s S<br />

In order to study the <strong>in</strong>tegral J E , we shall use polar coord<strong>in</strong>ates <strong>in</strong> the same way and with<br />

the same notation as <strong>in</strong> the proof of Theorem 3.3.1. Let {e 1 , Je 1 , . . . , e s−1 , Je s−1 } be a J-basis<br />

of E ∩ T p L s and let {e s+1 , Je s+1 , . . . , e n−1 , Je n−1 } be a J-basis of E ∩ (T p L s ) ⊥ . With respect<br />

to this orthonormal basis of E<br />

tr(II| E ) =<br />

n−1<br />

∑<br />

i=1,i≠s<br />

(k n (e i ) + k n (Je i )).<br />

If we denote by {f 1 , . . . , f 2n−1 } a basis of pr<strong>in</strong>cipal directions of S at p, we obta<strong>in</strong><br />

k n (e i ) =<br />

j=1<br />

2n−1<br />

∑<br />

j=1<br />

k n (f j )〈e i , f j 〉 2 =<br />

2n−1<br />

∑<br />

and us<strong>in</strong>g polar coord<strong>in</strong>ates with respect to JN, we get<br />

⎛<br />

⎞<br />

2n−1<br />

∑<br />

n−1<br />

∑<br />

tr(II| E ) = k j<br />

⎝ (〈e i , f j 〉 2 + 〈Je i , f j 〉 2 ) ⎠<br />

=<br />

2n−1<br />

∑<br />

j=1<br />

k j<br />

⎛<br />

⎝<br />

i=1,i≠s<br />

n−1<br />

∑<br />

i=1,i≠s<br />

j=1<br />

k j 〈e i , f j 〉 2 ,<br />

⎞<br />

(cos 2 (e i , JN, f j ) + cos 2 (Je i , JN, f j ) ⎠ .<br />

We denote by (u, v, w) the spherical angle with vertex v and sides on u and w, cos φ ij =<br />

cos(e i , JN, f j ) and cos φ ij<br />

= cos(Je i , JN, f j ). Then, to study the <strong>in</strong>tegral J E we have to deal<br />

with the follow<strong>in</strong>g <strong>in</strong>tegral<br />

∫S 2n−3 ∫ π/2<br />

0<br />

cos 2s−1 θ 1<br />

s<strong>in</strong> 2s−2 θ 1<br />

s<strong>in</strong> 2 α j s<strong>in</strong> 2n−3 θ 1·<br />

· (cos 2 φ 1j + · · · + cos 2 φ s−1,j + cos 2 φ s+1,j + · · · + cos 2 φ n−1,j )dθ 1 dS 2n−3<br />

= s<strong>in</strong> 2 Γ(s)Γ(n − s)<br />

α j ·<br />

2Γ(n)<br />

∫<br />

· (cos 2 φ 1j + · · · + cos 2 φ s−1,j + cos 2 φ s+1,j + · · · + cos 2 φ n−1,j )dS 2n−3<br />

S 2n−3<br />

where θ 1 and α i are def<strong>in</strong>ed <strong>in</strong> (3.10) and (3.11).


3.3 Mean curvature <strong>in</strong>tegral 55<br />

Denote by ˜S 2n−1 the subset of S 2n−1 def<strong>in</strong>ed by all po<strong>in</strong>ts not <strong>in</strong> span{N, JN}. Consider<br />

the well-def<strong>in</strong>ed map<br />

Π : ˜S2n−1 −→ {N, JN} ⊥<br />

v ↦→ proj {N,JN} ⊥ (v)<br />

||proj {N,JN} ⊥(v)||<br />

Note that Π(e a ) = e a , with a ∈ {1, 1, . . . , s − 1, s − 1, s + 1, s + 1, . . . , n − 1, n − 1}<br />

Then, V ⊥ <strong>in</strong>side {N, JN} ⊥ is<br />

E ⊥ ∩ {N, JN} ⊥ = ({N, JN, e s , Je s } ⊥ ) ⊥ ∩ {N, JN} ⊥ = {Π(e s ), JΠ(e s )}. (3.12)<br />

As cos(φ aj ) = cos(e a , JN, f j ) denotes the cos<strong>in</strong>e of the spherical angle with vertex JN and<br />

po<strong>in</strong>ts <strong>in</strong> e a and f j , by def<strong>in</strong>ition, it co<strong>in</strong>cides with 〈Π(e a ), Π(f j )〉 = 〈e a , Π(f j )〉. Then, as Π(f j )<br />

is a unit vector conta<strong>in</strong>ed <strong>in</strong> the vector sub<strong>space</strong> with basis {e 1 , Je 1 , . . . , e s−1 , Je s−1 , Π(e s ), JΠ(e s )}<br />

it is satisfied<br />

1 = 〈Π(f j ), Π(f j )〉 2<br />

= 〈e 1 , Π(f j )〉 2 + · · · + 〈Je s−1 , Π(f j )〉 2 + 〈Π(e s ), Π(f j )〉 2 + 〈JΠ(e s ), Π(f j )〉 2<br />

.<br />

and we get<br />

∫<br />

(cos 2 φ 1j + · · · + cos 2 φ s−1,j + cos 2 φ s+1,j + ... + cos 2 φ n−1,j )dS<br />

S∫<br />

2n−3<br />

= (1 − 〈Π(e s ), Π(f j )〉 2 − 〈JΠ(e s ), Π(f j )〉 2 )dS.<br />

S 2n−3<br />

Now, we use polar coord<strong>in</strong>ates θ 2 , θ 3 with respect to Π(f j ) such that<br />

〈Π(e s ), Π(f j )〉 = cos θ 2 , θ 2 ∈ (0, π),<br />

and<br />

〈JΠ(e s ), Π(f j )〉 = s<strong>in</strong>(Π(e s ), Π(f j )) cos(Π(e s ), Π(f j ), JΠ(f j )) = s<strong>in</strong> θ 2 cos θ 3 , θ 3 ∈ (0, π).<br />

By Lemma 3.1.2 and the relation<br />

π<br />

O 2n−3 = O 2n−5<br />

n − 2<br />

we have<br />

∫ π ∫ π<br />

∫S 2n−5<br />

0<br />

= O 2n−3 −<br />

(1 − cos 2 θ 2 − s<strong>in</strong> 2 θ 2 cos 2 θ 3 ) s<strong>in</strong> 2n−4 θ 2 s<strong>in</strong> 2n−5 θ 3 dθ 3 dθ 2 dS 1<br />

0<br />

∫ π ∫S 2n−4<br />

0<br />

∫ π<br />

cos 2 θ 2 s<strong>in</strong> 2n−4 θ 2 −<br />

∫S 2n−5<br />

= O 2n−3 − O √ √<br />

2n−3<br />

πΓ(n − 2) πΓ(n −<br />

1<br />

2n − 2 − O 2n−52<br />

4Γ(n − 1 2 ) Γ(n)<br />

( π<br />

= O 2n−5<br />

n − 2 − π<br />

2(n − 1)(n − 2) − π<br />

2(n − 1)(n − 2)<br />

O 2n−5 π<br />

=<br />

(2n − 4)<br />

2(n − 1)(n − 2)<br />

= O 2n−5π<br />

2(n − 1) .<br />

0<br />

∫ π<br />

cos 2 θ 3 s<strong>in</strong> 2n−5 θ 3 s<strong>in</strong> 2n−2 θ 2<br />

2 )<br />

)<br />

0


56 Average of the mean curvature <strong>in</strong>tegral<br />

Thus,<br />

J E = O 2n−3vol(G C (<br />

n−2,s−1 )n(s − 1) n −1 (<br />

∫ )<br />

(2n − 1)M 1 (S) − k n (JN)<br />

2s(2s − 1)(n − s)(n − 1) s)<br />

S<br />

and add<strong>in</strong>g both expressions of J E and J s we get the result<br />

J E + J s = O 2n−3vol(G C n−2,s−1 ) ( ) n −1<br />

·<br />

4s(2s − 1)(n − 1) s<br />

(<br />

n(s − 1)<br />

· ((2n − 1) + (2n − 1)<br />

n − s )M 1(S) + ( 2sn − n − s<br />

)<br />

n(s − 1)<br />

−<br />

n − s n − s<br />

∫S<br />

) k n (JN)<br />

= O 2n−3vol(G C n−2,s−1 ) ( ) n −1<br />

·<br />

4s(2s − 1)(n − 1) s<br />

( ∫ )<br />

2n − 1<br />

·<br />

n − s (n − s + ns − n)M 2sn − n − s − ns + n<br />

1(S) + k n (JN)<br />

n − s<br />

S<br />

= O 2n−3vol(G C n−2,s−1 ) ( ) n −1 ( ∫ )<br />

s(2n − 1)(n − 1)<br />

M 1 (S) + k n (JN) .<br />

4s(2s − 1)(n − 1) s<br />

n − s<br />

S<br />

It is natural to ask which functionals we have to <strong>in</strong>tegrate over the <strong>space</strong> of <strong>complex</strong> s-planes<br />

to obta<strong>in</strong> the mean curvature <strong>in</strong>tegral of the <strong>in</strong>itial hypersurface.<br />

Theorem 3.3.3. Let S ⊂ CK n (ɛ) be a compact (possibly with boundary) oriented hypersurface<br />

of class C 2 . If we def<strong>in</strong>e<br />

ν(S) = (2ns − n − s)M 1 (S) − n − s ∫<br />

k n (JN)dx,<br />

2s − 1<br />

S<br />

then<br />

∫<br />

L C r<br />

ν(S ∩ L r )dL r = ω 2n−2vol(G C n−2,s−1 )2(n − 1)(2n − 1)(s − 1)<br />

M 1 (S).<br />

(2s − 1)<br />

Proof. The result follows straightforward from Theorems 3.3.1 and 3.3.2.<br />

3.4 Reproductive valuations<br />

Def<strong>in</strong>ition 3.4.1. Suppose given, for each s ∈ N, a valuation <strong>in</strong> CK n (ɛ), ϕ (s) . It is said that<br />

the collection {ϕ (s) } satisfies the reproductive property if for any regular doma<strong>in</strong> Ω,<br />

∫<br />

ϕ (s) (Ω ∩ L s )dL s = c n,s ϕ(Ω),<br />

for some constant c n,s depend<strong>in</strong>g on n and s.<br />

L C s<br />

Remark 3.4.2. Recall that mean curvature <strong>in</strong>tegral for regular doma<strong>in</strong>s extend to all K(C n ).<br />

Also ∫ ∂Ω k n(JN) extends to K(C n ) s<strong>in</strong>ce it co<strong>in</strong>cides with Γ 2n−2,n−1 (Ω) (cf. Example 2.4.20).<br />

As neither the mean curvature <strong>in</strong>tegral M (s)<br />

1 (∂Ω ∩ L s), nor the <strong>in</strong>tegral of the normal<br />

curvature, ∫ ∂Ω∩L s<br />

k n (JÑ)dx, satisfy the reproductive property, it is natural to ask whether<br />

there exists some l<strong>in</strong>ear comb<strong>in</strong>ation of these such that satisfies this property. We consider a<br />

l<strong>in</strong>ear comb<strong>in</strong>ation s<strong>in</strong>ce, <strong>in</strong> C n , they constitute a basis of Val U(n)<br />

n−2 (Cn ).


3.4 Reproductive valuations 57<br />

Theorem 3.4.3. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>, s ∈ {1, ..., n−1}. Consider the smooth<br />

valuations def<strong>in</strong>ed by<br />

∫<br />

ϕ 1 (Ω) = M 1 (∂Ω) − k n (JN)<br />

∂Ω<br />

and<br />

Then,<br />

and<br />

∫<br />

L C s<br />

∫<br />

ϕ 2 (Ω) = (2s − 1)(2n − 1)M 1 (∂Ω) + k n (JN).<br />

∂Ω<br />

ϕ 1 (Ω ∩ L s )dL s = ω 2n−2vol(G C ( )<br />

n−2,s−1 )(s − 1)(2n − 1) n −1<br />

ϕ 1 (Ω)<br />

(2s − 1)(n − s)<br />

s<br />

∫<br />

L C s<br />

ϕ 2 (Ω ∩ L s )dL s = ω 2n−2vol(G C n−2,s−1 )<br />

(2s − 1)<br />

( ) n − 2 −1<br />

ϕ 2 (Ω).<br />

s − 1<br />

In C n , each of ϕ 1 (Ω) and ϕ 2 (Ω) expands a 1-dimensional sub<strong>space</strong> of reproductive valuations<br />

of degree 2n − 2.<br />

Proof. Let<br />

∫<br />

ν(Ω) = aM 1 (∂Ω) + b k n (JN).<br />

∂Ω<br />

We look for relations between a and b to be ν a reproductive valuation, that is,<br />

∫<br />

ν(Ω ∩ L s )dL s = λν(Ω).<br />

L C s<br />

From Theorems 3.3.1 and 3.3.2 we have<br />

∫<br />

∫ (<br />

)<br />

ν(Ω ∩ L s )dL s = aM 1 (∂Ω ∩ L s ) + b k n<br />

∫∂Ω∩L (JÑ) s<br />

L C s<br />

L C s<br />

= ω 2n−2vol(G C n−2,s−1 ) ( ) n −1((<br />

b(2s − 1)(2sn − n − s)<br />

) ∫<br />

a + k n (JN)+<br />

2s(2s − 1) s<br />

n − s<br />

∂Ω<br />

+ ( a(2ns − n − s)<br />

+ b(2s − 1) ) )<br />

(2n − 1)M 1 (∂Ω) .<br />

n − s<br />

Thus, ν is reproductive if and only if for some λ ∈ R<br />

( )<br />

a(2ns − n − s)<br />

(2n − 1)<br />

+ b(2s − 1) = λa,<br />

n − s<br />

b(2s − 1)(2sn − n − s)<br />

a + = λb.<br />

n − s<br />

Solv<strong>in</strong>g this system we get two solutions, a = −b, λ =<br />

1)(2n − 1), λ =<br />

2n(n − 1)<br />

.<br />

n − s<br />

2s(s − 1)(2n − 1)<br />

n − s<br />

and a = b(2s −<br />

Remark 3.4.4. Last theorem gives all valuations <strong>in</strong> Val U(n)<br />

n−2 (Cn ) such that they satisfy the<br />

reproductive property. Why are these valuations reproductive Are they special <strong>in</strong> some<br />

sense It shall be <strong>in</strong>terest<strong>in</strong>g to know the answer and also to have a geometric <strong>in</strong>terpretation<br />

for these valuations.


58 Average of the mean curvature <strong>in</strong>tegral<br />

3.5 Relation with some valuations def<strong>in</strong>ed by Alesker<br />

In Section 2.3 we recalled the def<strong>in</strong>ition of valuations U k,p <strong>in</strong> C n . They constitute a basis for<br />

Val U(n) (C n ). From this basis it can be established the follow<strong>in</strong>g theorem by Alesker<br />

Theorem 3.5.1. (Theorem 3.1.2 [Ale03]) Let Ω be a regular doma<strong>in</strong> <strong>in</strong> C n . Let 0 < q <<br />

n, 0 < 2p < k < 2q. Then,<br />

∫<br />

L C q<br />

U k,p (Ω ∩ L q )dL q =<br />

where constants γ p depend only on n, q and p.<br />

[k/2]+n−q<br />

∑<br />

p=0<br />

γ p · U k+2(n−q),p (Ω),<br />

In the follow<strong>in</strong>g theorem we give the constants γ p with arbitrary n, q and k = 2q − 2.<br />

Theorem 3.5.2. Let Ω be a regular doma<strong>in</strong> and 0 < q < n. Then,<br />

∫<br />

U 2q−2,p (Ω ∩ L q )dL q = ω 2q−2ω 2n−2 vol(G C n−2,q−1 )vol(GC q−2,q−p−1 )<br />

L C (q − p)(2q − 2p − 1) ( )(<br />

n−2 q−2<br />

) ·<br />

q<br />

q−1 q−p−1<br />

( )<br />

(2n − 3)(n − 1)(n − q + p)<br />

·<br />

U 2n−2,1 (Ω) − (2n − 1)n(n − q + p − 1)U 2n−2,0 (Ω) .<br />

ω 2n−2<br />

First, we express ∫ ∂Ω k n(JN) (a translation <strong>in</strong>variant cont<strong>in</strong>uous valuation) <strong>in</strong> terms of<br />

{U k,p }.<br />

Proposition 3.5.3. Let Ω be a regular doma<strong>in</strong> <strong>in</strong> C n . Then<br />

∫<br />

k n (JN)dp = n(2n − 3)(2n − 2)2 ω 2<br />

U 2n−2,1 (Ω) − 2n(2n − 1)(2n 2 − 4n + 1)ω 2 U 2n−2,0 (Ω).<br />

ω 2n−2<br />

∂Ω<br />

Proof. From the relations among valuations {U k,p } and mean curvature <strong>in</strong>tegrals (see (2.3))<br />

we have<br />

U 2n−2,0 (Ω) = 1 M 1 (∂Ω) (3.13)<br />

2nω 2<br />

and<br />

∫<br />

1<br />

U 2n−2,1 (Ω) =<br />

M 1 (∂Ω ∩ L n−1 )dL n−1 .<br />

(2n − 2)ω 2<br />

L C n−1<br />

Us<strong>in</strong>g Proposition 3.3.2 with s = n − 1, we get the result.<br />

Proof of the theorem 3.5.2. From the def<strong>in</strong>ition of U k,p we have<br />

∫<br />

1<br />

U k,p (Ω) =<br />

M k−2p (Ω ∩ L n−p )dL n−p<br />

2(n − p)ω 2n−k<br />

and from Theorem 3.3.2<br />

∫<br />

1<br />

U 2q−2,p (Ω ∩ L q ) =<br />

2(q − p)ω 2<br />

L C q−p<br />

G C n,n−p<br />

M 1 ((∂Ω ∩ L q ) ∩ L q−p )dL q−p<br />

) −1·<br />

O 2q−3 vol(G C q−2,q−p−1<br />

=<br />

) ( q<br />

8(q − p) 2 (q − 1)(2q − 2p − 1)ω 2 q − p<br />

(<br />

)<br />

2q(q − p) − 2q + p<br />

· (2q − 1) M 1 (∂Ω ∩ L q ) + k n<br />

p<br />

∫∂Ω∩L (JÑ) q<br />

where Ñ denotes the normal <strong>in</strong>ward vector field to ∂Ω ∩ L q as a hypersurface <strong>in</strong> L q . Us<strong>in</strong>g<br />

aga<strong>in</strong> Theorems 3.3.1 and 3.3.2, we express the <strong>in</strong>tegrals over L C q as an <strong>in</strong>tegral oer ∂Ω. F<strong>in</strong>ally,<br />

from the relation <strong>in</strong> Proposition 3.5.3 and (3.13) we get the result.


3.6 Example: sphere <strong>in</strong> CK 3 (ɛ) 59<br />

3.6 Example: sphere <strong>in</strong> CK 3 (ɛ)<br />

In this section we check Theorem 3.3.2 <strong>in</strong> the case of a sphere of radius R <strong>in</strong> CK 3 (ɛ).<br />

On page 18 we give an expression for the pr<strong>in</strong>cipal curvatures of a sphere of radius R <strong>in</strong><br />

CK n (ɛ). Us<strong>in</strong>g this expression we get<br />

∫<br />

and<br />

k n (JN) = 2 cot ɛ (2R)V (∂B R ) = 2 cos2 ɛ(R) + s<strong>in</strong> 2 ɛ(R)<br />

∂B R<br />

2 s<strong>in</strong> ɛ (R) cos ɛ (R)<br />

π 3<br />

3! 6 s<strong>in</strong>5 ɛ(R) cos ɛ (R)<br />

= π 3 (cos 2 ɛ(R) + s<strong>in</strong> 2 ɛ(R)) s<strong>in</strong> 4 ɛ(R) = π 3 (2 s<strong>in</strong> 6 ɛ(R) + s<strong>in</strong> 4 ɛ(R))<br />

= π 3 (2 cos 6 ɛ(R) − 5 cos 4 ɛ(R) + 4 cos 2 ɛ(R) − 1)<br />

M 1 (∂B R ) = π3 6<br />

3!4 (5 s<strong>in</strong>4 ɛ(R) cos 2 ɛ(R) + s<strong>in</strong> 6 ɛ(R)) = π3<br />

4 (6 s<strong>in</strong>6 ɛ(R) + 5 s<strong>in</strong> 4 ɛ(R))<br />

= π 3 ( 6 5 cos6 ɛ(R) − 13<br />

5 cos4 ɛ(R) + 8 5 cos2 ɛ(R) − 1 5 ).<br />

Thus, the right hand side of Theorem 3.3.2 is<br />

O 3 π 3<br />

144 ((42 + 2) cos6 ɛ(R) − (13 · 7 + 5) cos 4 ɛ(R) + (56 + 4) cos 2 ɛ(R) − (7 + 1))<br />

( 11<br />

= 2π 5 36 cos6 ɛ(R) − 2 3 cos4 ɛ(R) + 5<br />

12 cos2 ɛ(R) − 1 )<br />

.<br />

18<br />

The left hand side of Theorem 3.3.2 is<br />

∫<br />

M (2)<br />

1 (∂B R ∩ L 2 )dL 2 .<br />

L C 2<br />

Let us compute first M (2)<br />

1 (∂B R ∩ L 2 ), for a fixed <strong>complex</strong> 2-plane, L 2 . Recall that the <strong>in</strong>tersection<br />

between a sphere and L 2 is a sphere <strong>in</strong>side L 2 with radius r satisfy<strong>in</strong>g cos ɛ (R) =<br />

cos ɛ (r) cos ɛ (ρ) where ρ is the distance from the orig<strong>in</strong> of the sphere B R to the plane L 2 (cf.<br />

[Gol99, Lemma 3.2.13]).<br />

Thus,<br />

∫<br />

L C 2<br />

= 1 12<br />

M (2)<br />

1 (∂B R ∩ L 2 ) = 1 ∫<br />

(2 cot ɛ (r) + 2 cot ɛ (2r))<br />

3 ∂B R<br />

M (2)<br />

1 (∂B R ∩ L 2 )dL 2<br />

∫<br />

G C 3,1<br />

∫S 1 ∫ R<br />

= 2πV (GC 3,1 )2<br />

12<br />

= πV (GC 3,1 )<br />

3<br />

0<br />

∫ R<br />

0<br />

= 2 3 (cot ɛ(r) + cot ɛ (2r)) 4π2 s<strong>in</strong> 3<br />

2!<br />

ɛ(r) cos ɛ (r)<br />

= 1<br />

12 (4 cos4 ɛ(r) − 5 cos 2 ɛ(r) + 1)<br />

= 1 1<br />

12 cos 4 ɛ(R) (4 cos4 ɛ(R) − 5 cos 2 ɛ(R) cos 2 ɛ(ρ) + cos 4 ɛ(ρ)).<br />

cos 4 ɛ(ρ)<br />

cos 4 ɛ(ρ) (4 cos4 ɛ(R) − 5 cos 2 ɛ(R) cos 2 ɛ(ρ) + cos 4 ɛ(ρ))2 cos ɛ (ρ) s<strong>in</strong> ɛ (ρ)<br />

(4 cos 4 ɛ(R) − 5 cos 2 ɛ(R) cos 2 ɛ(ρ) + cos 4 ɛ(ρ))2 cos ɛ (ρ) s<strong>in</strong> ɛ (ρ)<br />

( 11<br />

12 cos6 ɛ(R) − 2 cos 4 ɛ(R) + 5 4 cos2 ɛ(R) − 1 6<br />

and we get the same result <strong>in</strong> both side of the expression <strong>in</strong> Theorem 3.3.2.<br />

)


Chapter 4<br />

Gauss-Bonnet Theorem and Crofton<br />

formulas for <strong>complex</strong> planes<br />

In this chapter we obta<strong>in</strong> an expression for the measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g a<br />

compact doma<strong>in</strong> <strong>in</strong> CK n (ɛ). That is, we give an expression of<br />

∫<br />

χ(Ω ∩ L r )dL r (4.1)<br />

L C r<br />

for a regular doma<strong>in</strong> Ω ⊂ CK n (ɛ) as a l<strong>in</strong>ear comb<strong>in</strong>ation of the so-called Hermitic <strong>in</strong>tr<strong>in</strong>sic<br />

volumes valuations <strong>in</strong> CK n (ɛ) (cf. Def<strong>in</strong>ition 2.4.11). The method we use consists on comput<strong>in</strong>g<br />

the variation, when the doma<strong>in</strong> moves along the flow <strong>in</strong>duced by a smooth vector field, of the<br />

measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g the convex doma<strong>in</strong>. From the theory of valuations on<br />

C n , we know that the expression is a l<strong>in</strong>ear comb<strong>in</strong>ation of certa<strong>in</strong> valuations. Thus, comput<strong>in</strong>g<br />

also the variation of these valuations and then compar<strong>in</strong>g both results we shall deduce the f<strong>in</strong>al<br />

expression.<br />

Us<strong>in</strong>g the same method we obta<strong>in</strong> an expression (cf. Theorem 4.4.1) for the Euler characteristic<br />

of a compact doma<strong>in</strong> <strong>in</strong> terms of its Gauss curvature of the boundary, its volume<br />

and others Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes. This expression is analogous to the one obta<strong>in</strong>ed <strong>in</strong><br />

[San04, page 309] for real <strong>space</strong> <strong>forms</strong>.<br />

Relat<strong>in</strong>g these two results we shall obta<strong>in</strong> another expression for the Euler characteristic.<br />

This one <strong>in</strong>volves the measure of <strong>complex</strong> hyperplanes <strong>in</strong>tersect<strong>in</strong>g the regular doma<strong>in</strong> (cf.<br />

Theorem 4.4.5).<br />

4.1 Variation of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes<br />

The study of the variation of a valuation when the doma<strong>in</strong> moves along the flow of a smooth<br />

vector field will be useful to deduce some properties of the valuation. In [BF08] it is given<br />

the variation of some valuations on C n and this variation is used to characterize monotone<br />

valuations. In this work, we give the variation of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes (cf. Def<strong>in</strong>ition<br />

2.4.11) on CK n (ɛ) and we use it to deduce expression (4.1) <strong>in</strong> terms of these valuations.<br />

In order to obta<strong>in</strong> the variation of Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes on CK n (ɛ) we follow the<br />

same method as <strong>in</strong> the proof of Corollary 2.6 <strong>in</strong> [BF08]. First, we recall the def<strong>in</strong>ition of the<br />

Rum<strong>in</strong> derivative, <strong>in</strong>troduced <strong>in</strong> [Rum94].<br />

Def<strong>in</strong>ition 4.1.1. Let µ ∈ Ω 2n−1 (S(CK n (ɛ))), let α be the contact form of S(CK n (ɛ)) and let<br />

α ∧ ξ ∈ Ω 2n−1 (S(CK n (ɛ))) be the unique (cf. [Rum94]) form such that d(µ + α ∧ ξ) is multiple<br />

of α. Then, the Rum<strong>in</strong> operator D is def<strong>in</strong>ed as<br />

Dµ := d(µ + α ∧ ξ).<br />

61


62 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

Let us recall the def<strong>in</strong>ition of the Reeb vector field over a contact manifold.<br />

Def<strong>in</strong>ition 4.1.2. Let M be a contact manifold with contact form α. The Reeb vector field<br />

T is the only vector field over M such that<br />

{<br />

iT α = 1,<br />

L T α = 0.<br />

(4.2)<br />

If the contact manifold is the fiber tangent bundle of a Riemann manifold, then the Reeb<br />

vector field co<strong>in</strong>cides with the geodesic flow (cf. [Bla76, page 17]). This is the situation <strong>in</strong> this<br />

work, we consider the unit tangent bundle of CK n (ɛ) (cf. Lemma 1.3.7 and Remark 1.3.8).<br />

Note that the condition L T α = 0 is equivalent to i T dα = 0. Indeed,<br />

L T α = i T dα + d(i T α) = i T dα = dα(T ).<br />

The follow<strong>in</strong>g lemma conta<strong>in</strong>s the value of the contraction of T with α, β, γ and θ i def<strong>in</strong>ed<br />

<strong>in</strong> Section 2.4.2.<br />

Lemma 4.1.3. In S(CK n (ɛ)) it is satisfied<br />

i T α = 1, i T θ 1 = γ,<br />

i T θ 2 = β, i T β = i T γ = i T θ 0 = i T θ s = 0.<br />

Proof. The first equality is a characterization of the Reeb vector field. Moreover, i T θ s =<br />

−i t dα = di T α − L T α = 0.<br />

As T is the geodesic flow, it satisfies α i (T ) = β i (T ) = 0 and α 1i = β 1i = 0, i ∈ {2, ..., n}.<br />

We get the result us<strong>in</strong>g Def<strong>in</strong>ition (2.6) extended to CK n (ɛ).<br />

In [BF08] it is proved the follow<strong>in</strong>g lemma, which allow to calculate the variation of a<br />

valuation def<strong>in</strong>ed from an <strong>in</strong>variant smooth form. The result is proved <strong>in</strong> C n but the same<br />

rema<strong>in</strong>s true for ɛ ≠ 0, and for any Riemann manifold. Here we repeat the proof <strong>in</strong> detail for<br />

CK n (ɛ).<br />

Lemma 4.1.4 ([BF08] Lemma 2.5). Suppose that Ω ⊂ CK n (ɛ) is a regular doma<strong>in</strong>, N the<br />

outward unit vector field to ∂Ω, X is a smooth vector field on CK n (ɛ) with flow F t and µ a<br />

smooth valuation given by a (2n − 1)-form ρ <strong>in</strong> S(CK n (ɛ)). Then<br />

∫<br />

d<br />

dt∣ µ(F t (Ω)) = δ X µ(Ω) =<br />

t=0<br />

N(Ω)<br />

〈X, N〉 i T (Dβ)<br />

where T is the Reeb vector field of S(CK n (ɛ)) and Dρ is the Rum<strong>in</strong> operator of ρ.<br />

Proof. Let ˜X be a lift of X at S(CK n (ɛ)) such that it preserves α, i.e. L ˜X(α) = 0. Then, if<br />

˜F denotes the flow of ˜X


4.1 Variation of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes 63<br />

δ X µ(Ω) = d ( ∫ )<br />

dt∣ β = d ( ∫ )<br />

∗ t=0 N(F t(Ω)) dt∣ ˜F t β<br />

t=0 N(Ω)<br />

∫<br />

=<br />

(1)<br />

=<br />

(3)<br />

=<br />

(4)<br />

=<br />

(5)<br />

=<br />

(6)<br />

=<br />

(7)<br />

=<br />

L ˜Xβ<br />

N(Ω)<br />

∫<br />

∫<br />

∫<br />

∫<br />

∫<br />

∫<br />

∫<br />

=<br />

N(Ω)<br />

N(Ω)<br />

N(Ω)<br />

N(Ω)<br />

N(Ω)<br />

N(Ω)<br />

N(Ω)<br />

∫<br />

(2)<br />

i ˜Xdβ + d(i ˜Xβ) = i ˜Xdβ<br />

N(Ω)<br />

i ˜XDβ − i ˜Xd(α ∧ η)<br />

i ˜XDβ<br />

i ˜X(α ∧ ρ)<br />

(i ˜Xα)ρ<br />

α( ˜X)i T (α ∧ ρ)<br />

〈X, N〉i T Dβ<br />

First, note that we can change the variation of F t (Ω) <strong>in</strong> the first <strong>in</strong>tegral for the Lie derivative<br />

of the <strong>in</strong>tegrated form s<strong>in</strong>ce N(F t (Ω)) = ˜F t (N(Ω)).<br />

For (1) and (2), we use the follow<strong>in</strong>g property of the Lie derivative, L ˜Xβ<br />

= i ˜Xdβ + d(i ˜Xβ),<br />

and that the second term is an exact form, thus the <strong>in</strong>tegral vanishes.<br />

Equality (3) follows directly from the def<strong>in</strong>ition of the Rum<strong>in</strong> operator.<br />

For (4) we use<br />

i ˜Xd(α ∧ η) = L ˜X(α ∧ η) − d(i ˜X(α ∧ η))<br />

and that the second term is an exact form. The first term can be rewritten as<br />

L ˜X(α ∧ η) = (L ˜Xα) ∧ η + α ∧ L ˜Xη,<br />

and so, its <strong>in</strong>tegral vanishes s<strong>in</strong>ce ˜X preserves α, which vanishes over the normal fiber bundle<br />

(cf. Lemma 1.3.12).<br />

As the Rum<strong>in</strong> operator is, by def<strong>in</strong>ition, a 2n-form multiple of α, and it is def<strong>in</strong>ed on the<br />

normal fiber bundle, we get (5).<br />

For (6), us<strong>in</strong>g the notion of contraction we get<br />

i ˜X(α ∧ ρ) = (i ˜Xα) ∧ ρ + α ∧ (i ˜Xρ).<br />

The second term vanishes over N(Ω).<br />

To get equality (7), we repeat the same argument as <strong>in</strong> (4) <strong>in</strong> order to obta<strong>in</strong> the form<br />

α ∧ ρ, which is the Rum<strong>in</strong> operator of β.<br />

F<strong>in</strong>ally, we recall the def<strong>in</strong>ition of α and that the <strong>in</strong>tegral is over the unit fiber normal<br />

bundle, so that the po<strong>in</strong>ts are (x, N) with x ∈ ∂Ω and N the unit normal vector on ∂Ω at<br />

x.<br />

The previous lemma allows us to compute the variation of any valuation given by a form,<br />

once we know its Rum<strong>in</strong> operator. In this chapter we give the variation of the Hermitian<br />

<strong>in</strong>tr<strong>in</strong>sic volumes <strong>in</strong> CK n (ɛ) (<strong>in</strong> [BF08] is given for ɛ = 0).<br />

In the follow<strong>in</strong>g lemma we give the derivative β k,q and γ k,q (def<strong>in</strong>ed <strong>in</strong> 2.4.6) us<strong>in</strong>g Lemma<br />

2.4.8. They will be used for the computation of the Rum<strong>in</strong> operator.


64 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

Lemma 4.1.5. In CK n (ɛ)<br />

and<br />

dβ k,q = c n,k,q (θ n−k+q<br />

0 ∧ θ k−2q<br />

1 ∧ θ q 2 − ɛ(n − k + q)α ∧ β ∧ θn−k+q−1 0 ∧ θ k−2q<br />

1 ∧ θ q 2 )<br />

dγ k,q =c n,k,q (θ n−k+q<br />

0 ∧ θ k−2q<br />

1 ∧ θ q 2 − ɛθn−k+q−1 0 ∧ θ k−2q<br />

1 ∧ θ q+1<br />

2 − ɛα ∧ β ∧ θ n−k+q−1<br />

0 ∧ θ k−2q<br />

1 ∧ θ q 2<br />

(n − k + q − 1)<br />

− ɛ α ∧ γ ∧ θ n−k+q−2<br />

0 ∧ θ k−2q+1<br />

1 ∧ θ q 2<br />

2<br />

(n − k + q − 1)<br />

− ɛ β ∧ γ ∧ θ 01 ∧ θ n−k+q−2<br />

0 ∧ θ k−2q<br />

1 ∧ θ q 2<br />

2<br />

).<br />

From the previous lemma and follow<strong>in</strong>g the method used by [BF08] we can compute the<br />

variation of B k,q and Γ k,q <strong>in</strong> CK n (ɛ).<br />

Notation 4.1.6. We denote<br />

∫<br />

˜B k,q = ˜B k,q (Ω) :=<br />

∂Ω<br />

∫<br />

〈X, N〉β k,q and ˜Γk,q = ˜Γ k,q (Ω) :=<br />

∂Ω<br />

〈X, N〉γ k,q .<br />

Proposition 4.1.7. Let X be a smooth vector field def<strong>in</strong>ed on CK n (ɛ) and Ω ⊂ CK n (ɛ) be<br />

a regular doma<strong>in</strong>. The variation <strong>in</strong> CK n (ɛ) of valuations B k,q and Γ k,q with respect to X is<br />

given by<br />

and<br />

δ X B k,q (Ω) = 2c n,k,q (c −1<br />

n,k−1,q (k − 2q)2˜Γk−1,q − c −1<br />

n,k−1,q−1 (n + q − k)q˜Γ k−1,q−1<br />

+c −1<br />

n,k−1,q−1 (n + q − k + 1 2 )q ˜B k−1,q−1 − c −1<br />

n,k−1,q (k − 2q)(k − 2q − 1) ˜B k−1,q<br />

+ɛ(c −1<br />

n,k+1,q+1 (k − 2q)(k − 2q − 1) ˜B k+1,q+1 − c −1<br />

n,k+1,q (n − k + q)(q + 1 2 ) ˜B k+1,q ))<br />

δ X Γ k,q (Ω) = 2c n,k,q<br />

(<br />

c −1<br />

n,k−1,q (k − 2q)2˜Γk−1,q − c −1<br />

n,k−1,q−1 (n + q − k)q˜Γ k−1,q−1<br />

+c −1<br />

n,k−1,q−1 (n + q − k + 1 2 )q ˜B k−1,q−1 − c −1<br />

n,k−1,q (k − 2q)(k − 2q − 1) ˜B k−1,q<br />

+ɛ ( c −1<br />

n,k+1,q+1 2(k − 2q)(k − 2q − 1) ˜B k+1,q+1 − c −1<br />

n,k+1,q ((n − k + q)(2q + 3 2 ) − 1 2 (q + 1)) ˜B k+1,q<br />

−c −1<br />

n,k+1,q+1 (k − 2q)2˜Γk+1,q+1 + c −1<br />

n,k+1,q (n − k + q − 1)(q + 1)˜Γ k+1,q<br />

−ɛ(c −1<br />

n,k+3,q+2 (k − 2q)(k − 2q − 1) ˜B k+3,q+2 − c −1<br />

n,k+3,q+1 (n − k + q − 1)(q + 3 2 ) ˜B k+3,q+1 ) )) .<br />

Proof. We first study the valuation given by β k,q .<br />

Lemma 4.1.4 provides an expression for the variation of a smooth valuation. In order to<br />

use this lemma, it is enough to f<strong>in</strong>d i T Dβ k,q and i T γ k,q modulo α and dα s<strong>in</strong>ce the latter <strong>forms</strong><br />

vanish over N(Ω) (cf. Lemma 1.3.12).<br />

We will use the follow<strong>in</strong>g fact from the proof of Proposition 4.6 <strong>in</strong> [BF08]: for max{0, k −<br />

n} ≤ q ≤ k/2 < n there exists an <strong>in</strong>variant form ξ k,q ∈ Ω 2n−1 (S(C n )) such that<br />

and<br />

dα ∧ ξ k,q ≡ −θ n−k+q<br />

0 θ k−2q<br />

1 θ q 2 mod(α), (4.3)<br />

ξ k,q ≡ βγθ n+q−k−1<br />

0 θ k−2q−2<br />

1 θ q−1<br />

2 (4.4)<br />

∧ ( (n + q − k)qθ1 2 )<br />

− (k − 2q)(k − 2q − 1)θ 0 θ 2 mod(α, dα).


4.1 Variation of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes 65<br />

In order to f<strong>in</strong>d δ X B k,q for general ɛ, we take a form ξ ɛ ∈ Ω 2n−1 (S(CK n (ɛ))) such that ξ ɛ (p,v) ≡<br />

ξ (p ′ ,v ′ ) when we identify T (p,v) S(CK n (ɛ)) and T (p ′ ,v ′ )C n , for every (p, v) ∈ S(CK n (ɛ)), (p ′ , v ′ ) ∈<br />

S(C n ). That is, as ξ ɛ is an <strong>in</strong>variant form, it can be expressed as a l<strong>in</strong>ear comb<strong>in</strong>ation of<br />

products with the <strong>forms</strong> α, β, θ 0 , θ 1 , θ 2 and θ s . We take this expression as a def<strong>in</strong>ition of ξ ɛ .<br />

From Lemma 4.1.5 and (4.3) we have that d(β k,q + c n,k,q α ∧ ξ ɛ ) ≡ 0 modulo α.<br />

By Lemma 2.4.8, the exterior differential of ξ ɛ is<br />

dξ ɛ ≡ θ n+q−k−1<br />

0 θ k−2q−2<br />

1 θ q−1<br />

2 ((n − k + q)qθ1 2 − (k − 2q)(k − 2q − 1)θ 0 θ 2 )<br />

∧ (γθ 1 − 2βθ 0 + 2ɛβθ 2 ) mod(α, dα)<br />

and the contraction of dβ k,q with respect to the vector field T , by Lemma 4.1.3, is<br />

i T dβ k,q ≡ c n,k,q θ n+q−k−1<br />

0 θ k−2q−1<br />

1 θ q−1<br />

2<br />

∧ ((k − 2q)γθ 0 θ 2 + qβθ 0 θ 1 − ɛ(n − k + q)βθ 1 θ 2 )<br />

mod(α).<br />

By substitut<strong>in</strong>g the last expressions <strong>in</strong> i T Dβ k,q ≡ i T dβ k,q − c n,k,q dξ ( mod α, dα), we get<br />

the result.<br />

The variation of Γ k,q with k ≠ 2q can be obta<strong>in</strong>ed us<strong>in</strong>g the relation among Γ k,q and B k,q<br />

given <strong>in</strong> Proposition 2.4.8 and the variation of B k,q .<br />

To compute δ X Γ 2q,q , note that dγ 2q,q has 3 terms not multiple of α (cf. Lemma 4.1.5). As<br />

before we take ξ ɛ 1 , ξɛ 2 ∈ Ω2n−1 (S(CK n (ɛ))) correspond<strong>in</strong>g to ξ 2q,q , and ξ 2q+2,q+1 respectively.<br />

Let us consider also<br />

ξ3 ɛ = n − q − 1 βγθ n−q−2<br />

0 θ q 2<br />

2<br />

. (4.5)<br />

Then the Rum<strong>in</strong> differential of γ 2q,q is given by Dγ 2q,q = d(γ 2q,q + c n,2q,q α ∧ (ξ1 ɛ − ɛξɛ 2 − ɛξɛ 3 )).<br />

Indeed, dα ∧ ξ1 ɛ cancels the first term of dγ 2q,q modulo α, and dα ∧ ξ2 ɛ cancels the second one.<br />

The third term is canceled exactly by dα ∧ ξ3 ɛ.<br />

Now, us<strong>in</strong>g Lemmas 4.1.5 and 4.1.3 we get<br />

i T dγ 2q,q ≡qβθ n−q<br />

0 θ q−1<br />

2 − ɛ(q + 2)βθ n−q−1<br />

and from (4.4) and (4.5)<br />

0 θ q 2 − ɛn − q − 1<br />

2<br />

γθ n−q−2<br />

0 θ 1 θ q 2 mod(α, dα),<br />

dξ ɛ 1 ≡ (n − q)qθ n−q−1<br />

0 θ q−1<br />

2 (γθ 1 − 2βθ 0 + 2ɛβθ 2 ) mod(α, dα).<br />

dξ ɛ 2 ≡ (n − q − 1)(q + 1)θ n−q−2<br />

0 θ q 2 (γθ 1 − 2βθ 0 + 2ɛβθ 2 ) mod(α, dα).<br />

dξ3 ɛ ≡ n − q − 1 θ n−q−2<br />

0 θ q 2<br />

2<br />

(γθ 1 − 2βθ 0 + 2ɛβθ 2 ) mod(α, dα).<br />

Plugg<strong>in</strong>g this <strong>in</strong>to i T Dγ 2q,q ≡ i T dγ 2q,q − c n,2q,q (dξ1 ɛ − ɛdξɛ 2 − ɛdξɛ 3 ) mod (α, dα) gives the result.<br />

Remark 4.1.8. For ɛ = 0 the variation of B k,q co<strong>in</strong>cides with the variation of Γ k,q and we get<br />

the result of Proposition 4.6 <strong>in</strong> [BF08].<br />

From the previous proposition we can obta<strong>in</strong> easily the variation of the Gauss curvature<br />

<strong>in</strong>tegral. We know that this variation vanishes <strong>in</strong> C n , for the Gauss-Bonnet theorem, but not<br />

<strong>in</strong> the other <strong>complex</strong> <strong>space</strong> <strong>forms</strong>.<br />

Corollary 4.1.9. In CK n (ɛ) the variation of the Gauss curvature <strong>in</strong>tegral is<br />

δ X M 2n−1 (∂Ω) = 2ɛω 2n−1 (2(n − 1)˜Γ 1,0 − (3n − 1) ˜B 1,0 + 3<br />

2π ɛ(2n − 1) ˜B 3,1 ).


66 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

Proof. First, we relate the Gauss curvature <strong>in</strong>tegral with Γ 0,0 (Ω) from Example 2.4.20.1<br />

M 2n−1 (∂Ω) = 2c−1 n,0,0<br />

(n − 1)! Γ 0,0(Ω) = 2nω 2n Γ 0,0 (Ω). (4.6)<br />

Thus, from Proposition 4.1.7 and us<strong>in</strong>g that c −1<br />

n,1,0 = (n − 1)!ω 2n−1, c −1<br />

n,3,1 = (n − 2)!ω 2n−3 and<br />

ω 2n−1 = (2n − 1)ω 2n−3 /2π we obta<strong>in</strong> the result:<br />

δ X M 2n−1 (∂Ω) =<br />

2ɛ<br />

(n − 1)! (−c−1 n,1,0 (3n − 1) ˜B 1,0 + 2c −1<br />

n,1,0 (n − 1)˜Γ 1,0 + c −1<br />

n,3,1 3ɛ(n − 1) ˜B 3,1 )<br />

= 2ɛ(−ω 2n−1 (3n − 1) ˜B 1,0 + 2ω 2n−1 (n − 1)˜Γ 1,0 + 3ɛω 2n−3 ˜B3,1 )<br />

= 2ɛω 2n−1 (−(3n − 1) ˜B 1,0 + 2(n − 1)˜Γ 1,0 + 3<br />

2π ɛ(2n − 1) ˜B 3,1 ).<br />

4.2 Variation of the measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g<br />

a regular doma<strong>in</strong><br />

Proposition 4.2.1. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>, X a smooth vector field on CK n (ɛ)<br />

with flow φ t and Ω t = φ t (Ω). Then<br />

∫<br />

∫<br />

( ∫ )<br />

d<br />

dt∣ χ(Ω t ∩ L r )dL r = 〈∂φ/∂t, N〉 σ 2r (II| V )dV dx<br />

t=0 G C n,r (Dp)<br />

L C r<br />

∂Ω<br />

where N is the outward normal field, D is the tangent distribution to ∂Ω and orthogonal to<br />

JN and σ 2r (II| V ) denotes the 2r-th symmetric elementary function of II restricted to V ∈<br />

G C n−1,r (D p).<br />

Proof. The proof is similar to the one <strong>in</strong> [Sol06, Theorem 4] for real <strong>space</strong> <strong>forms</strong>.<br />

Denote G C n−1,r (D) = {(p, l) | l ⊆ T p∂Ω, dim R l = 2r and Jl = l} = ⋃ p∈∂Ω GC n−1,r (D p).<br />

For each V ∈ G C n−1,r (D p), we take the parallel translation V t of V along φ t (x). Recall that<br />

parallel translation preserves the <strong>complex</strong> structure (cf. [O’N83, page 326]). Then we project<br />

V t orthogonally to D φt(x), obta<strong>in</strong><strong>in</strong>g a <strong>complex</strong> r-plane V ′<br />

t (at least for small values of t). We<br />

def<strong>in</strong>e<br />

γ : G C n−1,r (D) × (−ɛ, ɛ) −→ LC r<br />

((x, V ), t) ↦→ exp φt(x) V t ′ .<br />

Proposition 3 <strong>in</strong> [Sol06] rema<strong>in</strong>s true, without change, <strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong>. From this<br />

proposition and us<strong>in</strong>g a similar argument as <strong>in</strong> [Sol06, teorema 4] we get<br />

∫<br />

∫<br />

d<br />

1<br />

∑ ∂φ<br />

dt∣ χ(Ω t ∩ L r )dL r = lim<br />

sign〈<br />

t=0 L C h→0 h<br />

r<br />

G C n−1,r (D)×(0,h) ∂t , N〉 sign(σ 2r(II| V ))γ ∗ dL r<br />

∫<br />

= 〈 ∂φ<br />

G C n−1,r (D) ∂t , N〉 sign(σ 2r(II| V ))γ0(ι ∗ dφ∂t dL r )<br />

where the sum on the second <strong>in</strong>tegral runs over the tangencies of L r with the hypersurfaces<br />

∂Ω t with 0 < t < h.<br />

Consider a J-mov<strong>in</strong>g frame {g; g 1 , Jg 1 , ..., g n , Jg n } such that g((p, l), t) = φ(p, t), γ =<br />

〈g; g 1 , Jg 1 , ..., g r , Jg r 〉 ∩ CK n (ɛ) and Jg n ((p, l), t) = N t (outward unit vector to ∂Ω t at φ(p, t)).<br />

We may assume that the mov<strong>in</strong>g frame is def<strong>in</strong>ed <strong>in</strong> a neighborhood of L C r s<strong>in</strong>ce we are only<br />

<strong>in</strong>terested <strong>in</strong> regular po<strong>in</strong>ts of γ.<br />

(4.7)


4.2 Variation of the measure of <strong>complex</strong> r-planes <strong>in</strong>tersect<strong>in</strong>g a doma<strong>in</strong> 67<br />

Considerem the curve L r (t) given by the parallel translation of L r along the geodesic<br />

given by N, the outward normal vector to ∂Ω 0 . Recall that parallel translations preserves the<br />

<strong>complex</strong> structure (cf. [O’N83, page 326]). If P ∈ T Lr L C r denotes the tangent vector to L r (t)<br />

at t = 0, then<br />

ω i (P ) = 〈dg(P ), g i 〉 = 〈 d dt g(L r(t)), g i 〉 = 0, i ∈ {r + 1, r + 1, ..., n − 1, n − 1},<br />

ω n (P ) = 〈dg(P ), N〉 = 1, (4.8)<br />

ω ij (P ) = 〈∇g i (P ), g j 〉 = 〈 D dt g i(L r (t)), g j 〉 = 0,<br />

j ∈ {r + 1, r + 1, ..., n, n}, i ∈ {1, 1, ..., r, r}.<br />

The measure of <strong>complex</strong> r-planes <strong>in</strong> CK n (ɛ) is (cf. Proposition 1.5.5)<br />

From (4.8), we get<br />

dL r =<br />

∣<br />

n∧<br />

i=r+1<br />

ω i ∧ ω i<br />

∧<br />

i=1,...,r<br />

j=r+1,...,n<br />

dL r = |ω n |ι P dL r<br />

s<strong>in</strong>ce ι P dL r = | ∧ n−1<br />

h=r+1 ω h ∧ ω h ∧ ω n<br />

∧<br />

ωij |.Thus,<br />

with<br />

and we get<br />

ω ij ω ij<br />

∣ ∣∣∣∣∣∣∣<br />

.<br />

ι dγ∂t dL r = |ω n (dγ∂t)|ι P dL r + |ω n |ι dγ∂t ι P dL r<br />

ω n (dγ∂t) = 〈dg(dγ∂t), N〉 = 〈 ∂φ<br />

∂t , N〉,<br />

γ ∗ 0(ω n )(v) = 〈dg(dγ 0 (v)), N〉 = 0 ∀v ∈ T (p,l) G C n−1,r(T p ∂Ω 0 ),<br />

γ ∗ 0(ι dγ∂t dL r ) = |〈 ∂φ<br />

∂t , N〉|γ∗ 0(ι P dL r ).<br />

F<strong>in</strong>ally, us<strong>in</strong>g that ψ ∗ 0 (ι P dL r ) = |σ 2r (II| V )|dV dx, we get the result.<br />

Remark 4.2.2. The <strong>in</strong>tegral<br />

∫<br />

G C n,r<br />

σ 2r (II| V )dV (4.9)<br />

seems difficult to compute directly. However, we will f<strong>in</strong>d it by an <strong>in</strong>direct method. Recall that<br />

the analogous <strong>in</strong>tegral <strong>in</strong> real <strong>space</strong> <strong>forms</strong> is a multiple of an elementary symmetric function<br />

of the pr<strong>in</strong>cipal curvatures.<br />

For r = n − 1, the <strong>in</strong>tegral (4.9) can be easily computed <strong>in</strong> CK n (ɛ).<br />

Corollary 4.2.3. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>, X a smooth vector field on CK n (ɛ)<br />

with flow φ t , and Ω t = φ t (Ω). Then,<br />

∫<br />

d<br />

dt∣ t=0<br />

L C n−1<br />

χ(L n−1 ∩ Ω t )dL n−1 = ω 2n−1 ˜B1,0 (Ω).


68 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

Proof. From Proposition 4.2.1 we get the result s<strong>in</strong>ce there is just one <strong>complex</strong> hyperplane<br />

tangent to a po<strong>in</strong>t <strong>in</strong> ∂Ω. Thus,<br />

∫<br />

∫<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L n−1 )dL n−1 = 〈∂φ/∂t, N〉 σ 2n−2 (II| V )dV dx<br />

t=0 L C n−1<br />

∂Ω 0 G C n−1,n−1<br />

∫<br />

= 〈∂φ/∂t, N〉σ 2n−2 (II| D )dx<br />

∂Ω 0<br />

= 〈∂φ/∂t, N〉<br />

∫∂Ω β ∧ θn−1 0<br />

0<br />

(n − 1)!<br />

= c−1 n,1,0<br />

(n − 1)! ˜B 1,0 (Ω)<br />

= ω 2n−1 ˜B1,0 (Ω).<br />

4.3 Measure of <strong>complex</strong> r-planes meet<strong>in</strong>g a regular doma<strong>in</strong><br />

4.3.1 In the standard Hermitian <strong>space</strong><br />

Us<strong>in</strong>g that the measure of <strong>complex</strong> r-planes <strong>in</strong> C n meet<strong>in</strong>g a regular doma<strong>in</strong> is a l<strong>in</strong>ear comb<strong>in</strong>ation<br />

of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes, and Propositions 4.2.1 and 4.1.7 we f<strong>in</strong>d explicitly<br />

the coefficients of this l<strong>in</strong>ear comb<strong>in</strong>ation.<br />

Theorem 4.3.1. Let Ω ⊂ C n be a convex doma<strong>in</strong>, X a smooth vector field over C n , φ t the<br />

flow associated to X and Ω t = φ t (Ω). Then<br />

∫<br />

( )<br />

d<br />

n − 1 −1 ( ) n<br />

dt∣ χ(Ω t ∩ L r )dL r = vol(G C n−1,r)ω 2r+1 (r + 1)<br />

−1·<br />

t=0 L C r r<br />

r<br />

⎛<br />

⎞<br />

n−r−1<br />

∑<br />

( )<br />

· ⎝<br />

2n − 2r − 2q − 1 1<br />

n − r − q 4 ˜B n−r−q−1 2n−2r−1,q (Ω) ⎠ , (4.10)<br />

and<br />

∫<br />

L C r<br />

q=max{0,n−2r−1}<br />

( ) n − 1 −1 ( n<br />

χ(Ω ∩ L r )dL r = vol(G C n−1,r)ω 2r<br />

r r<br />

⎛<br />

· ⎝<br />

n−r<br />

∑<br />

q=max{0,n−2r}<br />

) −1·<br />

( )<br />

1 2n − 2r − 2q<br />

4 n−r−q n − r − q<br />

Proof. In order to simplify the follow<strong>in</strong>g computations, we consider<br />

and<br />

B k,q ′ = B′ k,q (Ω) := c−1<br />

n,k,q µ k,q(Ω),<br />

˜B ′ k,q = c−1 n,k,q ˜B k,q ,<br />

⎞<br />

µ 2n−2r,q (Ω) ⎠ . (4.11)<br />

Γ ′ 2q,q = Γ ′ 2q,q(Ω) := 2c −1<br />

n,2q,q µ 2q,q(Ω). (4.12)<br />

˜Γ′ k,q = 2c −1<br />

n,k,q ˜Γ k,q . (4.13)<br />

The functional ∫ L C r χ(Ω ∩ L r)dL r is a valuation on C n with degree of homogeneity 2n −<br />

2r. Thus, it can be expressed as a l<strong>in</strong>ear comb<strong>in</strong>ation of the elements of degree 2n − 2r,<br />

{µ 2n−2r,q | max{0, n − 2r} ≤ q ≤ n − r} (cf. Def<strong>in</strong>ition 2.4.11). Then, by Remark 2.4.12 and<br />

(2.8), we have<br />

∫<br />

L C r<br />

χ(Ω ∩ L r )dL r =<br />

n−r−1<br />

∑<br />

q=max{0,n−2r}<br />

C q B ′ 2n−2r,q + DΓ ′ 2n−2r,n−r (4.14)


4.3 Measure of <strong>complex</strong> r-planes meet<strong>in</strong>g a regular doma<strong>in</strong> 69<br />

for certa<strong>in</strong> constants C q , D which we wish to determ<strong>in</strong>e. This will be done by compar<strong>in</strong>g the<br />

variation of both sides of this equality.<br />

From here on we assume 2r < n. The case 2r ≥ n can be treated <strong>in</strong> the same way (cf.<br />

Remark 4.3.2).<br />

By Proposition 4.1.7, the variation of the right hand side of (4.14) is a l<strong>in</strong>ear comb<strong>in</strong>ation<br />

of the follow<strong>in</strong>g type<br />

n−r−1<br />

∑<br />

q=n−2r−1<br />

c q ˜B′ 2n−2r−1,q +<br />

n−r−1<br />

∑<br />

q=n−2r<br />

d q ˜Γ′ 2n−2r−1,q (4.15)<br />

where the coefficients c q and d q can be expressed <strong>in</strong> terms of a l<strong>in</strong>ear comb<strong>in</strong>ation with known<br />

coefficients of the variables C q and D, that still rema<strong>in</strong> unknown.<br />

The variation of the left hand side of (4.14), by Proposition 4.2.1 is<br />

∫<br />

∫<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L r )dL r = 〈∂φ/∂t, N〉 σ 2r (II| V )dV dx. (4.16)<br />

t=0<br />

L C r<br />

∂Ω<br />

G C n−1,r<br />

From Lemma 2.4.18 when pull<strong>in</strong>g-back the form γ k,q from N(Ω) to ∂Ω, one gets a polynomial<br />

expression P k,q of degree 2n − k − 1 <strong>in</strong> the coefficients h ij of II with i, j ∈ {1, 2, 2, . . . , n, n}.<br />

Moreover, for each q the monomials <strong>in</strong> P k,q conta<strong>in</strong><strong>in</strong>g only entries of the form h ii conta<strong>in</strong> the<br />

factor h 11 = II(JN, JN) and do not appear <strong>in</strong> any other P k,q ′ with q ′ ≠ q. Therefore, every<br />

non-trivial l<strong>in</strong>ear comb<strong>in</strong>ation of {P k,q } q must conta<strong>in</strong> the variable h 11 . On the other hand,<br />

the <strong>in</strong>tegral ∫ G<br />

σ C 2r (II| V )dV is a polynomial of the second fundamental form II restricted<br />

n−1,r<br />

to the distribution D = 〈N, JN〉 ⊥ , hence a polynomial not <strong>in</strong>volv<strong>in</strong>g h 11 . Compar<strong>in</strong>g the<br />

expressions of (4.15) and (4.16), it follows that d q = 0 for all q ∈ {n − 2r, . . . , n − r − 1}.<br />

As c q and d q depend on C q and D, we will obta<strong>in</strong> the value of c q once we know the value<br />

of C q and D. We will get their value from the equalities {d q = 0}. Note that this gives r<br />

equations, s<strong>in</strong>ce q runs from n − 2r to n − r − 1 <strong>in</strong> (4.15). As for the unknowns, we need to<br />

f<strong>in</strong>d r constants C q plus the constant D <strong>in</strong> (4.14).<br />

We will get an extra equation by tak<strong>in</strong>g II| D = Id and equat<strong>in</strong>g (4.16) to (4.15). Then,<br />

for any pair (n, r) we have a compatible l<strong>in</strong>ear system s<strong>in</strong>ce constants <strong>in</strong> (4.14) exist. Next we<br />

f<strong>in</strong>d the solution.<br />

Let us relate explicitly the coefficients {c q } and {d q } <strong>in</strong> (4.15) with C q and D <strong>in</strong> (4.14).<br />

To simplify the range of the subscripts, we denote d n−r−a with a = 1, . . . , r and c n−r−a with<br />

a = 1, . . . , r + 1.<br />

Coefficient d n−r−1 . From the variation of B ′ k,q <strong>in</strong> Cn (Proposition 4.1.7), the coefficient of<br />

˜Γ ′ 2n−2r−1,n−r−1 comes from the variation of B′ 2n−2r,n−r−1 and Γ′ 2n−2r,n−r . Then,<br />

d n−r−1 = −2r(n − r)D + (2n − 2r − 2(n − r − 1)) 2 C n−r−1<br />

= 4C n−r−1 − 2r(n − r)D. (4.17)<br />

Coefficient d n−r−a , a = 2, . . . , r. The coefficient of ˜Γ ′ 2n−2r−1,n−r−a comes from the variation<br />

of B 2n−2r,n−r−a ′ and B′ 2n−2r,n−r−a+1 . Then,<br />

d n−r−a = (2n − 2r − 2(n − r − a)) 2 C n−r−a − (2r + n − r − a + 1 − n)(n − r − a + 1)C n−r−a+1<br />

= 4a 2 C n−r−a − (r − a + 1)(n − r − a + 1)C n−r−a+1 . (4.18)<br />

Coefficient c n−r−1 . The coefficient of ˜B ′ 2n−2r−1,n−r−1 comes from the variation of B′ 2n−2r,n−r−1<br />

and Γ ′ 2n−2r,n−r . Then,<br />

c n−r−1 = 4(r + 1/2)(n − r)D − 4C n−r−1<br />

= 2(2r + 1)(n − r)D − 4C n−r−1 . (4.19)


70 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

Coefficient c n−r−a , a = 2, . . . , r − 2. The coefficient of ˜B′ 2n−2r−1,n−r−a comes from the<br />

variation of B ′ 2n−2r,n−r−a and B′ 2n−2r,n−r−a+1 . Then,<br />

c n−r−a = −2(2a)(2a − 1)C n−r−a + 2(r − a + 3/2)(n − r − a + 1)C n−r−a+1<br />

= −4a(2a − 1)C n−r−a + (2r − 2a + 3)(n − r − a + 1)C n−r−a+1 . (4.20)<br />

Coefficient c n−2r−1 . The coefficient of ˜B ′ 2n−2r−1,n−2r−1 comes from the variation of B′ 2n−2r,n−2r .<br />

Then,<br />

c n−2r−1 = (2r − 2(r + 1) + 3)(n − r − (r + 1) + 1)C n−2r<br />

= (n − 2r)C n−2r . (4.21)<br />

Now, we solve the l<strong>in</strong>ear system given by {d n−r−a = 0} where a ∈ {1, . . . , r}. From<br />

equations (4.17) and (4.18) the system is given by<br />

{ r(n − r)D = 2Cn−r−1<br />

4a 2 C n−r−a = (n − r − a + 1)(r − a + 1)C n−r−a+1 .<br />

Thus,<br />

(n − r − a + 1) · · · · · (n − r)(r − a + 1) · · · · r<br />

C n−r−a =<br />

2 · 4 a−1 a 2 (a − 1) 2 · · · · · 1 2 D<br />

(n − r)!r!<br />

=<br />

2 2a−1 (n − r − a)!(r − a)!a!a! D<br />

= D ( )( n − r r<br />

2 2a−1 . (4.22)<br />

a a)<br />

To obta<strong>in</strong> the value of D, we compute ∫ G<br />

σ C 2r (p)dV and β 2n−2r−1,n−r−a ′ <strong>in</strong> case II| D(p) =<br />

n−1,r<br />

λId for λ > 0, which occurs when Ω is a geodesic ball. On one hand, we have<br />

∫<br />

σ 2r (p)(λId| V )dV = λ 2r vol(G C n−1,r).<br />

G C n−1,r<br />

On the other hand, if II| D = λId, then the connection <strong>forms</strong> satisfy α 1i = λω i and β 1i = λω i .<br />

Thus, θ 1 = 2λθ 2 and θ 0 = λ 2 θ 2 and we obta<strong>in</strong><br />

β ′ 2n−2r−1,n−r−a(p) = λ 2r (β ∧ θ r−a+1<br />

0 ∧ θ 2a−2<br />

1 ∧ θ n−r−a<br />

2 )(p)<br />

= 2 2a−2 λ 2r (β ∧ θ n−1<br />

2 )(p) = 2 2a−2 λ 2r (n − 1)!.<br />

So, the equation<br />

∑r+1<br />

vol(G C n−1,r) = c n−r−a 2 2a−2 (n − 1)!<br />

a=1<br />

must be satisfied.


4.3 Measure of <strong>complex</strong> r-planes meet<strong>in</strong>g a regular doma<strong>in</strong> 71<br />

Substitut<strong>in</strong>g equations (4.19), (4.20) and (4.21) <strong>in</strong> the last equation gives<br />

vol(G C n−1,r )<br />

(n − 1)!<br />

= (2(2r + 1)(n − r)D − 4C n−r−1 )<br />

r∑<br />

+ 2 2a−2 ((2r − 2a + 3)(n − r + a + 1)C n−r−a+1 − 4a(2a − 1)C n−r−a )<br />

a=2<br />

Thus,<br />

+ 2 2r (n − 2r)C n−2r<br />

= 2(2r + 1)(n − r)D + 4((2r − 1)(n − r − 1) − 1)C n−r−1<br />

∑r−1<br />

+ (−2 2a−2 4a(2a − 1) + 2 2a (2r − 2a + 1)(n − r − a))C n−r−a<br />

a=2<br />

+ (2 2r (n − 2r) − 2 2r−2 4r(2r − 1))C n−2r<br />

r∑<br />

= 2(2r + 1)(n − r)D + 2 2a ((2r − 2a + 1)(n − r − a) − a(2a − 1))C n−r−a<br />

(<br />

(4.22)<br />

= D 2(n − r)!r!<br />

2 n!<br />

= D<br />

r!(n − r − 1)!<br />

r∑<br />

a=0<br />

a=1<br />

D = vol(GC n−1,r )<br />

2 n!<br />

C n−r−a = vol(GC n−1,r )<br />

4 a n!<br />

(2r − 2a + 1)(n − r − a) − a(2a − 1)<br />

(n − r − a)!(r − a)!a!a!<br />

( ) n − 1 −1<br />

,<br />

r<br />

( ) n − 1 −1 ( )( n − r r<br />

r a a)<br />

and, for 2r < n, we have<br />

∫<br />

r∑<br />

χ(Ω ∩ L r )dL r = C n−r−a B 2n−2r,n−r−a ′ + DΓ ′ 2n−2r,n−r<br />

L C r<br />

and<br />

∫<br />

d<br />

dt∣ t=0<br />

L C r<br />

a=1<br />

= vol(GC n−1,r )<br />

2 n!<br />

( ) (<br />

n − 1<br />

−1 r∑ ( n − r<br />

r<br />

a<br />

a=1<br />

χ(Ω t ∩ L r )dL r = (2(2r + 1)(n − r)D − 4C n−r−1 )B ′ 2n−2r−1,n−r−1<br />

+<br />

)<br />

)( r<br />

a)<br />

2 −2a+1 B ′ 2n−2r,n−r−a + Γ ′ 2n−2r,n−r<br />

r∑<br />

((2r − 2a + 3)(n − r + a + 1)C n−r−a+1 − 4a(2a − 1)C n−r−a )B 2n−2r−1,n−r−a<br />

′<br />

a=2<br />

+ (n − 2r)C n−2r B 2n−2r−1,n−2r−1<br />

′<br />

= vol(GC n−1,r ) ( ) (<br />

n − 1<br />

−1 ∑r+1<br />

( )( ) )<br />

n − r r + 1 a<br />

n! r<br />

a a 4 ˜B ′ a−1 2n−2r−1,n−r−a . (4.23)<br />

a=1<br />

F<strong>in</strong>ally, we use the relation <strong>in</strong> (4.12) and (2.8) to obta<strong>in</strong> the result.<br />

Remark 4.3.2. If 2r ≥ n, then formula (4.10) follows directly from the relations among the<br />

different bases of valuations on C n given <strong>in</strong> [BF08] and the follow<strong>in</strong>g relation <strong>in</strong> [Ale03]<br />

∫<br />

χ(Ω ∩ L r )dL r = 1 ∫<br />

M 2r−1 (∂Ω ∩ L r )dL r = cU<br />

O 2(n−r),n−r<br />

2r−1<br />

L C r<br />

L C r<br />

)


72 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

for a certa<strong>in</strong> constant c com<strong>in</strong>g from the different normalizations <strong>in</strong> dL r .<br />

Corollary 4.3.3. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>, X a smooth vector field over CK n (ɛ),<br />

φ t the flow associated to X and Ω t = φ t (Ω). Then<br />

∫<br />

( )<br />

d<br />

n − 1 −1 ( ) n<br />

dt∣ χ(Ω t ∩ L r )dL r = vol(G C n−1,r)ω 2r+1 (r + 1)<br />

−1·<br />

t=0 L C r r<br />

r<br />

⎛<br />

⎞<br />

n−r−1<br />

∑<br />

( )<br />

· ⎝<br />

2n − 2r − 2q − 1 1<br />

n − r − q 4 ˜B n−r−q−1 2n−2r−1,q (Ω) ⎠ . (4.24)<br />

q=max{0,n−2r−1}<br />

Proof. Compar<strong>in</strong>g equation (4.10) and Proposition 4.2.1 <strong>in</strong> case ɛ = 0 shows that if Ω is a<br />

regular convex doma<strong>in</strong>, then<br />

∫<br />

)<br />

〈X, N〉 σ 2r (II|V )dV dx<br />

∂Ω<br />

( ∫<br />

G C n,r<br />

equals the right hand side of equation above. By tak<strong>in</strong>g a vector field X that vanishes outside<br />

an arbitrarily small neighborhood of a fixed x ∈ ∂Ω, we deduce the follow<strong>in</strong>g equality between<br />

<strong>forms</strong><br />

(∫ )<br />

σ 2r (II| V )dV dx = ω 2r+1<br />

)( n<br />

G C n−1,r r)vol(G C n−1,r)(r + 1)· (4.25)<br />

(Tx∂Ω)<br />

·<br />

n−r−1<br />

∑<br />

q=max{0,n−2r−1}<br />

( n−1<br />

r<br />

( 2n − 2r − 2q − 1<br />

n − r − q<br />

)<br />

cn,2n−2r−1,n−r−q<br />

4 n−r−q−1 β ∧ θ r−q+1<br />

0 ∧ θ 2q−2<br />

1 ∧ θ n−r−q<br />

2 .<br />

This equation can be written as P (II)dx = Q(II)dx where P and Q are polyomials with entries<br />

<strong>in</strong> the second fundamental form. These polynomials concide for any positive def<strong>in</strong>ed bil<strong>in</strong>eal<br />

form. Thus, P = Q and (4.25) holds for regular doma<strong>in</strong>s (not necessarily convex doma<strong>in</strong>s).<br />

Moreover, it is valid <strong>in</strong> CK n (ɛ) for any ɛ. Apply<strong>in</strong>g Proposition 4.2.1 we get the result.<br />

Corollary 4.3.4. Equation (4.11) holds for any regular doma<strong>in</strong> not necessarily convex.<br />

Proof. Consider Ω t = φ t (Ω) with φ t a given flow.<br />

From the last corollary, it is known the variation of the left hand side of (4.11).<br />

By Proposition 4.1.7, the variation of the right hand side is a l<strong>in</strong>ear comb<strong>in</strong>ation of<br />

{B k,q , Γ k,q }. By Theorem 4.3.1 this l<strong>in</strong>ear comb<strong>in</strong>ation co<strong>in</strong>cides with the right hand side<br />

of (4.24).<br />

Thus, the variation of both sides of (4.11) co<strong>in</strong>cides. So, the difference between both<br />

members of (4.11) is constant.<br />

Take φ t such that φ t (Ω) converges to a po<strong>in</strong>t for t → ∞. Both sides of (4.11) tend to zero<br />

when t → ∞, thus their difference vanishes for all t.<br />

4.3.2 In <strong>complex</strong> <strong>space</strong> <strong>forms</strong><br />

Theorem 4.3.5. Let Ω be a regular doma<strong>in</strong> <strong>in</strong> CK n (ɛ). Then<br />

∫<br />

( ) n − 1<br />

χ(Ω ∩ L r )dL r = vol(G C n−1,r)<br />

−1· (4.26)<br />

L C r<br />

r<br />

⎛<br />

⎞<br />

n−1<br />

∑<br />

( ) n −1 ∑k−1<br />

( )<br />

· ( ɛ k−(n−r) ω 2n−2k<br />

⎝<br />

1 2k − 2q<br />

k<br />

4 k−q µ 2k,q (Ω) + (k + r − n + 1)µ 2k,k (Ω) ⎠<br />

k − q<br />

k=n−r<br />

+ ɛ r (r + 1)vol(Ω)).<br />

q=max{0,2k−n}


4.3 Measure of <strong>complex</strong> r-planes meet<strong>in</strong>g a regular doma<strong>in</strong> 73<br />

Proof. We will show that both sides have the same variation δ X with respect to any vector<br />

field X. This implies the result: one can take a deformation Ω t of Ω such that Ω t converges<br />

to a po<strong>in</strong>t. Then both sides of (4.26) have the same derivative, and both vanish <strong>in</strong> the limit.<br />

The variation of the left hand side of (4.26) is given by Corollary 4.3.3. The variation of<br />

the right hand side can be computed by us<strong>in</strong>g Proposition 4.1.7, and δ X V = 2 ˜B 2n−1,n−1 . In<br />

order to simplify the computations we rewrite the right hand side of (4.26) as<br />

+<br />

n−1<br />

∑<br />

j=n−r<br />

By Proposition 3.8 we have<br />

+<br />

n−1<br />

∑<br />

j=n−r<br />

C r (Ω) := vol(GC n−1,r )<br />

n!<br />

⎛<br />

ɛ j−n+r ⎝ j − n + r + 1 Γ ′ 2j,j +<br />

2<br />

δ X C r (Ω) = vol(GC n−1,r)<br />

n!<br />

( ) n − 1 −1<br />

{ɛ r (r + 1)n!V<br />

r<br />

j−1<br />

∑<br />

q=max(0,2j−n)<br />

( )<br />

1 n − j j<br />

4 j−q j − q)(<br />

q<br />

B ′ 2j,q<br />

⎞<br />

⎠}.<br />

( ) −1 n − 1<br />

[ɛ r n(r + 1)δ x ˜B′ 2n−1,n−1 (4.27)<br />

r<br />

ɛ j−n+r j − n + r + 1 {−2(n − j)j˜Γ ′ 2j−1,j−1 + 2ɛ(n − j − 1)(j + 1)˜Γ ′ 2j+1,j<br />

2<br />

+4(n − j + 1 (<br />

2 )j ˜B j + 1<br />

2j−1,j−1 ′ + 4ɛ − (n − j)(2j + 3 )<br />

2<br />

2 ) ˜B 2j+1,j ′ + 4ɛ 2 (n − j − 1)(j + 3 2 ) ˜B 2j+3,j+1}]<br />

′<br />

+<br />

n−1<br />

∑<br />

∑j−1<br />

j=n−r q=max{0,2j−n}<br />

ɛ j−n+r ( ) n − j j<br />

4 j−q {(2j − 2q)<br />

j − q)( 2˜Γ′ q<br />

2j−1,q<br />

−(n + q − 2j)q˜Γ ′ 2j−1,q−1 + 2(n + q − 2j + 1 2 )q ˜B ′ 2j−1,q−1 − 2(2j − 2q)(2j − 2q − 1) ˜B ′ 2j−1,q<br />

+2ɛ(2j − 2q)(2j − 2q − 1) ˜B ′ 2j+1,q+1 − 2ɛ(n − 2j + q)(q + 1 2 ) ˜B ′ 2j+1,q}.<br />

We will show that the previous expression is <strong>in</strong>dependent of ɛ; i.e. all the terms conta<strong>in</strong><strong>in</strong>g ɛ<br />

cancel out. Hence, δ X C r (Ω) co<strong>in</strong>cides with (4.24) s<strong>in</strong>ce we know this happens for ɛ = 0. This<br />

will f<strong>in</strong>ish the proof.<br />

We concentrate first on the terms with ˜B k,q ′ . By putt<strong>in</strong>g together similar terms, the third<br />

l<strong>in</strong>e of (4.27) is<br />

n−1<br />

∑<br />

h=n−r+1<br />

ɛ h−n+r 2{(h − n + r + 1)(n − h + 1 2 )h + (h − n + r)(h 2 − (n − h + 1)(2h − 1 )) (4.28)<br />

2<br />

+(h − n + r + 1)(n − h + 1)(h − 1 2 )} ˜B ′ 2h−1,h−1<br />

−ɛ r {(r + 2)n − 1} ˜B ′ 2n−1,n−1 + (2r + 1)(n − r) ˜B ′ 2n−2r−1,n−r−1.<br />

By putt<strong>in</strong>g together similar terms, the double sum <strong>in</strong> (4.27) (forgett<strong>in</strong>g for the moment<br />

the terms with ˜Γ ′ k,q ) becomes<br />

n−1<br />

∑<br />

h−2<br />

∑<br />

h=n−r a=max(−1,2h−n−1)<br />

−<br />

n−1<br />

∑<br />

h−1<br />

∑<br />

h=n−r a=max(0,2h−n)<br />

ɛ h−n+r ( )( )<br />

n − h h<br />

4 h−a−1 2(n + a − 2h + 3 h − a − 1 a + 1<br />

2 )(a + 1) ˜B 2h−1,a<br />

′<br />

ɛ h−n+r ( )<br />

n − h h<br />

4 h−a 2(2h − 2a)(2h − 2a − 1)<br />

h − a)(<br />

a<br />

˜B 2h−1,a<br />


74 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

−<br />

+<br />

n∑<br />

h−1<br />

∑<br />

h=n−r+1 a=max(1,2h−n−1)<br />

n∑<br />

h−2<br />

∑<br />

h=n−r+1 a=max(0,2h−n−2)<br />

ɛ h−n+r ( n − h + 1<br />

4 h−a h − a<br />

ɛ h−n+r ( n − h + 1<br />

4 h−a−1 h − a − 1<br />

)( ) h − 1<br />

2(2h − 2a)(2h − 2a − 1)<br />

a − 1<br />

˜B 2h−1,a<br />

′<br />

)( h − 1<br />

a<br />

)<br />

2(n − 2h + a + 2)(a + 1 2 ) ˜B ′ 2h−1,a.<br />

Note that the terms with a = −1 or a = 2h − n − 2 vanish, if they occur. Then, one checks<br />

that all the terms <strong>in</strong> the above expression cancel out except those with h = n − r, n, and those<br />

with a = h − 1. Clearly the terms correspond<strong>in</strong>g to h = n − r are <strong>in</strong>dependent of ɛ. The terms<br />

with h = n sum up ɛ r (n − 1) ˜B 2n−1,n−1 ′ , and together with the similar term appear<strong>in</strong>g <strong>in</strong> (4.28)<br />

cancel out the first term <strong>in</strong> (4.27). F<strong>in</strong>ally, the terms with a = h − 1 are cancelled with the<br />

sum <strong>in</strong> (4.28).<br />

With a similar but shorter analysis one checks that the multiples of ˜Γ ′ k,q<br />

cancel out completely.<br />

This shows that (4.27) is <strong>in</strong>dependent of ɛ, and f<strong>in</strong>ishes the proof.<br />

Remark 4.3.6. The coefficients of µ k,q and vol <strong>in</strong> (4.26) were found by solv<strong>in</strong>g a l<strong>in</strong>ear system<br />

of equations, which we write down <strong>in</strong> the appendix.<br />

4.4 Gauss-Bonnet formula <strong>in</strong> CK n (ɛ)<br />

Theorem 4.4.1. Let Ω be a regular doma<strong>in</strong> <strong>in</strong> CK n (ɛ). Then<br />

ω 2n χ(Ω) =(n + 1)ɛ n vol(Ω) + (4.29)<br />

⎛<br />

⎞<br />

n−1<br />

∑ (n − c)ω 2n−2c ɛ c<br />

+<br />

n ( ∑c−1<br />

( )<br />

) ⎝<br />

1 2c − 2q<br />

n−1<br />

4 c−q µ 2c,q (Ω) + (c + 1)µ 2c,c (Ω) ⎠ .<br />

c − q<br />

c<br />

c=0<br />

q=max{0,2c−n}<br />

Remark 4.4.2. For ɛ = 0 we have the Gauss-Bonnet formula <strong>in</strong> C n ∼ = R 2n , where it is known<br />

χ(Ω) = 1<br />

2nω 2n<br />

M 2n−1 (∂Ω) = µ 0,0 (Ω),<br />

which co<strong>in</strong>cides with the expression <strong>in</strong> the previous result.<br />

Here we prove the certa<strong>in</strong>ty of (4.29) but <strong>in</strong> the appendix we give a constructive proof of<br />

the result.<br />

Proof. We proceed analogously to the proof of Theorem 4.3.5. In fact, the same computations<br />

of the previous proof show (<strong>in</strong> case r = n) that the right hand side of (4.29) has null variation.<br />

For ɛ = 0 equation (4.29) is the well know Gauss-Bonnet formula <strong>in</strong> C n ∼ = R 2n . For ɛ ≠ 0,<br />

we take a smooth deformation of Ω to get a doma<strong>in</strong> conta<strong>in</strong>ed <strong>in</strong> a ball of radius r. Under<br />

this deformation, the right hand side of (4.29) rema<strong>in</strong>s constant. By tak<strong>in</strong>g r small enough,<br />

the difference between both sides can be made arbitarily small. Hence, they co<strong>in</strong>cide.<br />

Although <strong>in</strong> (4.29) does not appear the Gauss curvature, we can easily get the follow<strong>in</strong>g<br />

expression.<br />

Corollary 4.4.3. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>. Then,<br />

2nω 2n χ(Ω) =M 2n−1 (∂Ω) + 2n(n + 1)ɛ n vol(Ω)+<br />

⎛<br />

n−1<br />

∑<br />

( ) n −1<br />

+ 2nω 2n−2c ɛ c ⎝<br />

c<br />

c=1<br />

∑c−1<br />

q=max{0,2c−n}<br />

⎞<br />

( )<br />

1 2c − 2q<br />

4 c−q µ 2c,q (Ω) + (c + 1)µ 2c,c (Ω) ⎠ .<br />

c − q


4.4 Gauss-Bonnet formula <strong>in</strong> CK n (ɛ) 75<br />

Proof. Apply relation (4.6) <strong>in</strong> (4.29).<br />

Remarks 4.4.4. 1. The Gauss-Bonnet-Chern formula <strong>in</strong> <strong>space</strong>s of constant sectional curvature<br />

k and even dimension, for a regular doma<strong>in</strong> Ω is given by<br />

O 2m−1 χ(Ω) = M 2n−1 (∂Ω) + c n−3 M 2n−3 (∂Ω) + · · · + c 1 M 1 (∂Ω) + (|k|) n/2 vol(Ω),<br />

where c j are known and depend on the curvature k.<br />

Note that <strong>in</strong> the previous expression appear all mean curvature <strong>in</strong>tegrals with odd subscript<br />

and the volume. In formula (4.29) <strong>in</strong> CK n (ɛ), ɛ ≠ 0, also appear all the Hermitian<br />

<strong>in</strong>tr<strong>in</strong>sic volumes <strong>in</strong> CK n (ɛ) with the first subscript odd.<br />

2. In [Sol06] it is given an expression of the Gauss-Bonnet-Chern formula <strong>in</strong> <strong>space</strong> of constant<br />

sectional curvature k us<strong>in</strong>g the measure of planes of codimension 2 meet<strong>in</strong>g the<br />

doma<strong>in</strong>. The obta<strong>in</strong>ed formula for Ω ⊂ RK n (ɛ) is<br />

nω n χ(Ω) = M n−1 (∂Ω) +<br />

2k ∫<br />

χ(Ω ∩ L n−2 )dL n−2 . (4.30)<br />

ω n−1 L n−2<br />

A natural question is whether <strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong>, there exists a similar expression<br />

relat<strong>in</strong>g the Gauss curvature <strong>in</strong>tegral with the Euler characteristic and the measure of<br />

some <strong>complex</strong> planes meet<strong>in</strong>g the doma<strong>in</strong><br />

or<br />

c 0 χ(Ω) = n−1<br />

∑<br />

∫<br />

M 2n−1 (∂Ω) + c q<br />

q=1<br />

c 0 χ(Ω) = n−1<br />

∑<br />

n−1<br />

∑<br />

∫<br />

M 2n−1 (∂Ω) + M 2q+1 (∂Ω) + c q<br />

q=0<br />

L C q<br />

χ(Ω ∩ L q )dL q<br />

q=1<br />

L C q<br />

χ(Ω ∩ L q )dL q .<br />

Tak<strong>in</strong>g variation on both sides <strong>in</strong> these expressions, we get that these expressions cannot<br />

hold <strong>in</strong> general (for n = 2 and n = 3 we can choose constants satisfy<strong>in</strong>g them). Anyway,<br />

<strong>in</strong> Theorem 4.4.5 we give a similar expression. Perhaps, if we knew a formula for the<br />

measure of totally real planes meet<strong>in</strong>g a doma<strong>in</strong> we could f<strong>in</strong>d a more similar expression.<br />

3. For n = 2, the Gauss-Bonnet-Chern formula was already known <strong>in</strong> CK n (ɛ). It was given<br />

<strong>in</strong> [Par02]. From Theorem 4.4.1 we get the same expression, which can be written as<br />

χ(Ω) = 1 ( ( )<br />

)<br />

1 1<br />

π 2 2 Γ′ 0,0 + ɛ<br />

4 B′ 2,0 + Γ ′ 2,1 + 6ɛ 2 vol(Ω) .<br />

This expression can also be stated as<br />

χ(Ω) = 1 (<br />

2π 2 M 3 (∂Ω) + 3ɛ ( ∫<br />

M 1 (∂Ω) +<br />

2<br />

∂Ω<br />

)<br />

)<br />

k n (JN) + 12ɛ 2 vol(Ω)<br />

(4.31)<br />

and<br />

(<br />

χ(Ω) = 1<br />

∫<br />

2π 2 M 3 (∂Ω) + 2ɛ χ(∂Ω ∩ L 1 )dL 1 + ɛ ∫<br />

)<br />

k n (JN) + 12ɛ 2 vol(Ω) .<br />

L C 2<br />

1<br />

∂Ω<br />

In the follow<strong>in</strong>g result we express the Euler characteristic <strong>in</strong> terms of the Gauss curvature<br />

<strong>in</strong>tegral, the volume, the measure of <strong>complex</strong> hyperplanes meet<strong>in</strong>g a doma<strong>in</strong> and the valuations<br />

µ 2c,c . This formula generalizes (4.30) <strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong>.


76 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

Theorem 4.4.5. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong> CK n (ɛ). Then,<br />

∫<br />

ω 2n χ(Ω) = ɛ χ(Ω ∩ L n−1 )dL n−1 +<br />

L C n−1<br />

= 1<br />

∫<br />

2n M 2n−1(∂Ω) + ɛ<br />

L C n−1<br />

n∑<br />

c=0<br />

ɛ c ω 2n<br />

µ 2c,c (Ω)<br />

ω 2c<br />

χ(Ω ∩ L n−1 )dL n−1 +<br />

n∑<br />

c=1<br />

ɛ c ω 2n<br />

µ 2c,c (Ω).<br />

ω 2c<br />

Proof. From Theorems 4.3.5 and 4.4.1 we get the stated formula<br />

⎛<br />

⎞<br />

n−1<br />

∑ ɛ c n−1<br />

c! ∑<br />

( )<br />

χ(Ω) = ⎝<br />

1 2c − 2q<br />

π c<br />

4 c−q B 2c,q (Ω) + (c + 1)Γ 2c,c (Ω) ⎠ + ɛn (n + 1)!<br />

c − q<br />

π n vol(Ω)<br />

c=0 q=max{0,2c−n}<br />

⎛<br />

⎞<br />

n−1<br />

∑ ɛ c n−1<br />

c! ∑<br />

( )<br />

= Γ 0,0 (Ω)+ ⎝<br />

1 2c − 2q<br />

π c 4 c−q B 2c,c (Ω) + (c + 1)Γ 2c,c (Ω) ⎠+ ɛn (n + 1)!<br />

c − q<br />

π n vol(Ω)<br />

c=1 q=max{0,2c−n}<br />

⎛<br />

⎞<br />

= Γ 0,0 (Ω) + ɛ n! n−1<br />

∑ ɛ c−1 c!π n−c n−1<br />

∑<br />

( )<br />

⎝<br />

1 2c − 2q<br />

π n n!<br />

4 c−q B 2c,c (Ω) + cΓ 2c,c (Ω) ⎠ +<br />

c − q<br />

+ ɛ n! n−1<br />

∑<br />

π n<br />

c=1<br />

c=1<br />

= Γ 0,0 (Ω) + ɛ n!<br />

π n ∫<br />

ɛ c−1 c!π n−c<br />

n!<br />

L C n−1<br />

q=max{0,2c−n}<br />

Γ 2c,c (Ω) + ɛn (n + 1)!<br />

π n vol(Ω)<br />

n−1<br />

∑<br />

χ(Ω ∩ L n−1 )dL n−1 +<br />

c=1<br />

ɛ c (<br />

c!<br />

ɛ n )<br />

π c Γ (n + 1)!<br />

2c,c(Ω) +<br />

π n − ɛn n!n<br />

π n vol(Ω).<br />

4.5 Another method to compute the measure of <strong>complex</strong> l<strong>in</strong>es<br />

meet<strong>in</strong>g a regular doma<strong>in</strong><br />

From Theorem 4.3.5 we can give an expression of the measure of <strong>complex</strong> l<strong>in</strong>es meet<strong>in</strong>g a<br />

regular doma<strong>in</strong> (just tak<strong>in</strong>g r = 1). Here, we give another method to obta<strong>in</strong> this expression,<br />

us<strong>in</strong>g the results <strong>in</strong> Chapter 3.<br />

4.5.1 Measure of <strong>complex</strong> l<strong>in</strong>es meet<strong>in</strong>g a regular doma<strong>in</strong> <strong>in</strong> C n<br />

Proposition 4.5.1. Let Ω ⊂ C n be a regular doma<strong>in</strong>. Then,<br />

∫<br />

χ(Ω ∩ L 1 )dL 1 =<br />

ω (<br />

∫ )<br />

2n−4<br />

(2n − 1)M 1 (∂Ω) + k n (Jn) .<br />

4n(n − 1)<br />

∂Ω<br />

L C 1<br />

Proof. Recall that each <strong>complex</strong> l<strong>in</strong>e is isometric to C. Gauss-Bonnet formula <strong>in</strong> C n for<br />

hypersurfaces ∂Ω states<br />

M 2n−1 (∂Ω) = 2nω 2n χ(Ω).<br />

Apply<strong>in</strong>g Gauss-Bonnet formula <strong>in</strong> C and Proposition 3.3.2 with s = 1 we get the result<br />

∫<br />

χ(Ω ∩ L 1 )dL 1 = 1 ∫ ∫<br />

k g dpdL 1 = ω (<br />

∫ )<br />

2n−2<br />

(2n − 1)M 1 (∂Ω) + k n (Jn) .<br />

L C 2π<br />

1<br />

L C 1 ∂Ω∩L 1<br />

4nω 2 ∂Ω


4.5 Another method to compute the measure of <strong>complex</strong> l<strong>in</strong>es 77<br />

Although Gauss-Bonnet formula is known <strong>in</strong> C n for n ≥ 1, we cannot apply the same<br />

method to give the expression of the measure of s-planes meet<strong>in</strong>g a regular doma<strong>in</strong> s<strong>in</strong>ce<br />

the <strong>in</strong>tegral ∫ L C r M 2r−1(∂Ω ∩ L r )dL r is not <strong>in</strong> general known. In the next section we get<br />

an expression for this <strong>in</strong>tegral us<strong>in</strong>g the Gauss-Bonnet formula and the measure of <strong>complex</strong><br />

r-planes meet<strong>in</strong>g a regular doma<strong>in</strong>.<br />

4.5.2 Measure of <strong>complex</strong> l<strong>in</strong>es meet<strong>in</strong>g a regular doma<strong>in</strong> <strong>in</strong> CP n and CH n<br />

The follow<strong>in</strong>g result is given, for <strong>in</strong>stance, <strong>in</strong> [ÁPF04].<br />

Proposition 4.5.2 ([ÁPF04]). Let Ω be a regular doma<strong>in</strong> <strong>in</strong> CPn or CH n . Then,<br />

∫<br />

vol 2s (Ω ∩ L s )dL s = Cvol 2n (Ω).<br />

L C s<br />

The value of the constant C, it is not known, but now we shall need it explicitly.<br />

Proposition 4.5.3. Let Ω be a regular doma<strong>in</strong> <strong>in</strong> CP n or CH n . Then,<br />

∫<br />

vol 2s (Ω ∩ L s )dL s = vol(G C n,n−s)vol 2n (Ω).<br />

L C s<br />

Proof. In order to f<strong>in</strong>d C we apply last proposition to a ball of radius R. Let L s be a <strong>complex</strong><br />

s-plane meet<strong>in</strong>g B R at a distance ρ from the center of the ball. From Lemma 3.2.13 <strong>in</strong> [Gol99],<br />

we have that the <strong>in</strong>tersection B R ∩ L s is a ball of <strong>complex</strong> dimension s and radius r such that<br />

cos ɛ (R) = cos ɛ (r) cos ɛ (ρ).<br />

The expression of the volume of a geodesic ball of radius R <strong>in</strong> CK n (ɛ) is (cf. [Gra73])<br />

vol 2n (B R ) =<br />

πn<br />

|ɛ| n n! s<strong>in</strong>2n ɛ (R).<br />

Us<strong>in</strong>g this expression we get<br />

vol 2s (L s ∩ B R ) =<br />

( ) π s<br />

1<br />

|ɛ| s!<br />

( cos<br />

2<br />

ɛ (R) − cos 2 ) s<br />

ɛ(ρ)<br />

cos 2 .<br />

ɛ(ρ)<br />

On the other hand, the Jacobian of the change of variables to polar coord<strong>in</strong>ates is given by<br />

(cf. [Gra73])<br />

cos ɛ (R) s<strong>in</strong>ɛ<br />

2n−1 (R)<br />

|ɛ| n−1/2 .<br />

Then, us<strong>in</strong>g Proposition 1.5.8, we get<br />

∫<br />

L C s<br />

vol 2s (L s ∩ B R )dL s = πs vol(G C n,n−s)O 2(n−s)−1<br />

|ɛ| n−1/2 s!<br />

·<br />

∫ R<br />

0<br />

s∑<br />

( ) s<br />

(−1) i+1 ·<br />

i<br />

i=0<br />

s<strong>in</strong> 2(n−s)−1<br />

ɛ (ρ) cos 2i<br />

ɛ (R) cos 2(s−i)+1<br />

ɛ (ρ)dρ<br />

= πs vol(G C n,n−s)O 2(n−s)−1<br />

|ɛ| n−1/2 ·<br />

s!<br />

s∑ ∑s−i<br />

i∑<br />

( )( )( 2(n−s+k+j)<br />

s s − i i s<strong>in</strong><br />

· (−1) i+1 ɛ (R)<br />

√ .<br />

i j k)<br />

ɛ(n − s + j)<br />

i=0 j=0 k=0


78 Gauss-Bonnet Theorem and Crofton formulas for <strong>complex</strong> planes<br />

From Proposition 4.5.2, this expression is a multiple of<br />

vol 2n (B R ) =<br />

πn<br />

|ɛ| n n! s<strong>in</strong>2n ɛ (R).<br />

Thus, all terms <strong>in</strong> the sum are zero except for 2(n − s + k + j) = 2n, i.e. k + j = s, which<br />

together with j ≤ s − i, k ≤ i implies j = s − i, k = i, and we get<br />

∫<br />

vol 2s (L s ∩ B R )dL s<br />

L C s<br />

F<strong>in</strong>ally, from equality<br />

= πs vol(G C n,n−s)O 2(n−s)−1<br />

|ɛ| n s!<br />

= πs vol(G C n,n−s)O 2(n−s)−1<br />

2|ɛ| n<br />

s∑<br />

( ) s s<strong>in</strong><br />

(−1) i+1 2n<br />

ɛ (R)<br />

i 2(n − i)<br />

i=0<br />

s<strong>in</strong> 2n<br />

ɛ (R)<br />

s∑<br />

i=0<br />

= πs vol(G C n,n−s)O 2(n−s)−1 (n − s − 1)!<br />

2|ɛ| n n!<br />

and us<strong>in</strong>g O 2(n−s)−1 = 2(n − s)ω 2(n−s) = 2<br />

(−1) i+1<br />

i!(s − i)!(n − i)<br />

s<strong>in</strong> 2n<br />

ɛ (R).<br />

C<br />

πn<br />

|ɛ| n n! = πs vol(G C n,n−s)O 2(n−s)−1 (n − s − 1)!<br />

2|ɛ| n ,<br />

n!<br />

πn−s<br />

(n−s−1)!<br />

, we get the value of the constant C.<br />

Corollary 4.5.4. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>. Then,<br />

∫<br />

χ(Ω ∩ L 1 )dL 1 =<br />

ω (<br />

∫<br />

)<br />

2n−4<br />

(2n − 1)M 1 (∂Ω)+ k n (JN)+8nɛvol(Ω) .<br />

4n(n − 1)<br />

∂Ω<br />

L C 1<br />

where k n (JN) denotes the normal curvature <strong>in</strong> the direction JN.<br />

Proof. Us<strong>in</strong>g Gauss-Bonnet formula <strong>in</strong> H 2 (−4) we have (cf. [San04, page 309])<br />

∫<br />

χ(Ω ∩ L 1 )dL 1 = 1 ∫<br />

M 1 (∂Ω ∩ L 1 )dL 1 − 2 ∫<br />

vol(Ω ∩ L 1 )dL 1 ,<br />

2π L C π<br />

1<br />

L C 1<br />

and us<strong>in</strong>g Proposition 4.5.3 with s = 1, and Proposition 3.3.2 we get the result.<br />

Corollary 4.5.5. If Ω is a regular doma<strong>in</strong> <strong>in</strong> CK 2 (ɛ), ɛ ≠ 0, then<br />

∫<br />

Ω∩L 1 ≠∅<br />

χ(∂Ω ∩ L 1 )dL 1 = 1 4<br />

(M 1 (∂Ω) − 1 3ɛ M 3(∂Ω) + 4ɛvol(Ω) + 2π2<br />

3ɛ χ(Ω) )<br />

.<br />

Proof. From previous corollary, with n = 2, we have<br />

∫<br />

χ(Ω ∩ L 1 )dL 1 = 1 (<br />

∫<br />

)<br />

3M 1 (∂Ω) + k n (JN) + 16ɛvol(Ω) .<br />

L C 8<br />

1 ∂Ω<br />

Isolat<strong>in</strong>g ∫ ∂Ω k n(JN) <strong>in</strong> expression (4.31) we get the stated result.<br />

Note that the previous corollary cannot be extended to ɛ = 0 s<strong>in</strong>ce the expression (4.31)<br />

does not conta<strong>in</strong> the term ∫ ∂Ω k n(JN).<br />

L C 1


4.6 Total Gauss curvature <strong>in</strong>tegral C n 79<br />

4.6 Total Gauss curvature <strong>in</strong>tegral C n<br />

Theorem 4.6.1. If Ω ⊂ C n is a regular doma<strong>in</strong>, then<br />

∫<br />

L C r<br />

( ) n − 1 −1 ( ) n<br />

M 2r−1 (∂Ω∩L r )dL r = 2rω2rvol(G 2 C n−1,r)<br />

−1·<br />

r r<br />

⎛<br />

⎞<br />

n−r<br />

∑<br />

( )<br />

· ⎝<br />

1 2n − 2r − 2q<br />

4 n−r−q µ 2n−2r,q (Ω) ⎠ .<br />

n − r − q<br />

q=max{0,n−2r}<br />

Proof. On one hand, by Gauss-Bonnet formula <strong>in</strong> C n and the relation (4.6), we have<br />

∫<br />

∫<br />

∫<br />

M 2r−1 (∂Ω ∩ L r )dL r = 2rω r χ(Ω ∩ L r )dL r = 2rω 2r µ 0,0 (Ω ∩ L r )dL r . (4.32)<br />

L C r<br />

On the other hand, by Theorem 4.3.1, we have<br />

∫<br />

L C r<br />

L C r<br />

( ) n − 1 −1 ( n −1<br />

χ(∂Ω ∩ L r )dL r = vol(G C n−1,r)ω 2r<br />

r r)<br />

⎛<br />

⎞<br />

n−r<br />

∑<br />

( )<br />

· ⎝<br />

1 2n − 2r − 2q<br />

4 n−r−q µ 2n−2r,q (Ω) ⎠ .<br />

n − r − q<br />

q=max{0,n−2r}<br />

If we equate both expressions and we use the relation (4.6), we get the result.<br />

L C r


Chapter 5<br />

Other Crofton formulas<br />

In the previous chapter we give an expression for the measure of <strong>complex</strong> planes <strong>in</strong>tersect<strong>in</strong>g<br />

a regular doma<strong>in</strong> <strong>in</strong> a <strong>complex</strong> <strong>space</strong> form. Complex planes <strong>in</strong> CK n (ɛ) are totally geodesic<br />

submanifolds, but, by Theorem 1.4.6, there are other totally geodesic submanifolds. Totally<br />

real planes are also totally geodesic submanifolds <strong>in</strong> CK n (ɛ) for any ɛ (cf. Theorem 1.4.6).<br />

Moreover, for ɛ = 0, all submanifolds generated by the exponential map of a vector sub<strong>space</strong><br />

holomorphically isometric to C k ⊕ R k−2p are totally geodesic. Note that <strong>complex</strong> planes and<br />

totally real planes are particular cases of these submanifolds, for (k, p) = (2p, p) and (k, p) =<br />

(k, 0), respectively.<br />

In this chapter we obta<strong>in</strong> an expression for the measure of planes of type (2n−p, n−p), the<br />

so-called coisotropic planes, <strong>in</strong>tersect<strong>in</strong>g a doma<strong>in</strong> <strong>in</strong> C n , and an expression for the measure of<br />

Lagrangian planes <strong>in</strong> CK n (ɛ).<br />

5.1 Space of (k, p)-planes<br />

First, we recall the def<strong>in</strong>ition of (k, p)-plane <strong>in</strong> C n , as it is given <strong>in</strong> [BF08].<br />

Def<strong>in</strong>ition 5.1.1. Suppose that V is a real vector <strong>space</strong> and L n k<br />

(V ) denote the <strong>space</strong> of all<br />

aff<strong>in</strong>e sub<strong>space</strong>s of dimension k <strong>in</strong> V . If V = C n , considered as a real vector <strong>space</strong>, then the<br />

<strong>space</strong> of (k, p)-planes, L k,p (C n ) ⊂ L n k (Cn ) is def<strong>in</strong>ed as the subset of all sub<strong>space</strong> of (real)<br />

dimension k that can be expressed as the orthogonal direct sum of a <strong>complex</strong> sub<strong>space</strong> of<br />

<strong>complex</strong> dimension p and a totally real sub<strong>space</strong> of (real) dimension (k − 2p).<br />

We denote the elements of L k,p (C n ) by L k,p and the Grassmannian of all (k, p)-planes<br />

through the orig<strong>in</strong> <strong>in</strong> a vector <strong>space</strong> V by G n,k,p (V ).<br />

From the previous def<strong>in</strong>ition, L k,p (C n ) is the orbit of C p ⊕ R k−2p under the action of<br />

C n ⋊ U(n) (which are the holomorphic isometries of C n ).<br />

The notion of (k, p)-plane was extended to CH n . In [Gol99] and [Hsi98], they are def<strong>in</strong>ed<br />

as particular cases of the so-called l<strong>in</strong>ear submanifolds.<br />

Def<strong>in</strong>ition 5.1.2. The image of the exponential map from a po<strong>in</strong>t x ∈ CH n of a vector<br />

sub<strong>space</strong> <strong>in</strong> T x CH n is called l<strong>in</strong>ear submanifold.<br />

The image of the exponential map from a po<strong>in</strong>t x ∈ CH n of a (k, p)-plane <strong>in</strong> T x CH n is<br />

called l<strong>in</strong>ear (k, p)-plane.<br />

This def<strong>in</strong>ition could be also stated <strong>in</strong> CP n , but do<strong>in</strong>g so, we obta<strong>in</strong> submanifolds with<br />

s<strong>in</strong>gularities except when the submanifold is totally geodesic, i.e. for <strong>complex</strong> planes (which<br />

correspond to (2p, p)-planes), and for totally real planes (which correspond to (k, 0)-planes).<br />

In CH n , l<strong>in</strong>ear submanifolds are not always totally geodesic submanifolds. They are totally<br />

geodesic just for <strong>complex</strong> planes and totally real planes (cf. Theorem 1.4.6).<br />

81


82 Other Crofton formulas<br />

5.1.1 Bisectors<br />

As <strong>in</strong> <strong>complex</strong> hyperbolic <strong>space</strong> there are no totally geodesic real hypersurfaces, it is natural<br />

to look for real hypersurfaces with similar properties to those expected for a totally geodesic<br />

real hypersurface. In [Gol99, page 152] it is answered that they are the so-called bisectors, also<br />

denoted by sp<strong>in</strong>al superfaces.<br />

Def<strong>in</strong>ition 5.1.3. Let z 1 , z 2 be two different po<strong>in</strong>ts <strong>in</strong> CH n . The bisector equidistant from z 1<br />

and z 2 is def<strong>in</strong>ed as<br />

E(z 1 , z 2 ) = {z ∈ CH n | d(z 1 , z) = d(z 2 , z)}<br />

where d(z, z i ) denotes the distance between po<strong>in</strong>ts z i and z at CK n (ɛ) (see Proposition 1.2.3).<br />

Def<strong>in</strong>ition 5.1.4. Let z 1 , z 2 be two different po<strong>in</strong>ts <strong>in</strong> CH n .<br />

• The <strong>complex</strong> geodesic Σ def<strong>in</strong>ed by z 1 and z 2 is the <strong>complex</strong> sp<strong>in</strong>e of the bisector E(z 1 , z 2 ).<br />

• The real sp<strong>in</strong>e of the bisector E(z 1 , z 2 ), σ(z 1 , z 2 ) is the <strong>in</strong>tersection between the bisector<br />

and the <strong>complex</strong> sp<strong>in</strong>e, i.e.<br />

σ(z 1 , z 2 ) = E(z 1 , z 2 ) ∩ Σ(z 1 , z 2 ) = {z ∈ Σ | d(z 1 , z) = d(z 2 , z)}.<br />

• A slide of E is a <strong>complex</strong> hyperplane Π −1<br />

Σ (s) where Π : CHn −→ Σ denotes the orthogonal<br />

projection over Σ.<br />

Remark 5.1.5. 1. The set of all slides <strong>in</strong> a bisector def<strong>in</strong>es a foliation of the bisector by<br />

<strong>complex</strong> hyperplanes.<br />

2. The real sp<strong>in</strong>e is a (real) geodesic <strong>in</strong> CH n s<strong>in</strong>ce Σ is totally geodesic and isometric to<br />

H 2 , and <strong>in</strong> the real hyperbolic <strong>space</strong>, the bisector l<strong>in</strong>e of two given po<strong>in</strong>ts is a geodesic.<br />

3. Each geodesic γ ⊂ CH n is the real sp<strong>in</strong>e of a unique bisector. Indeed, take the <strong>complex</strong><br />

l<strong>in</strong>e Σ conta<strong>in</strong><strong>in</strong>g γ and the orthogonal projection Π Σ to Σ. Then, Π −1<br />

Σ<br />

(γ) def<strong>in</strong>es a<br />

bisector.<br />

Example 5.1.6. In CH 2 with the projective model, the bisector with respect to z 1 = [(1, 0, i)]<br />

and z 2 = [(1, 0, −i)] is<br />

E(z 1 , z 2 ) = {[(1, z, t)] ∈ CH n | z ∈ C, t ∈ R}.<br />

This expression can be obta<strong>in</strong>ed directly us<strong>in</strong>g the formula for the distance between 2 po<strong>in</strong>ts<br />

given at Proposition 1.2.3. {[(<br />

The <strong>complex</strong> sp<strong>in</strong>e is 1, 0, i λ − µ )]<br />

}<br />

with λ, µ ∈ C both nonzero and the real sp<strong>in</strong>e<br />

λ + µ<br />

is {[1, 0, t]}.<br />

Proposition 5.1.7. The isometries of CH n act transitively over the <strong>space</strong> of bisectors.<br />

Proof. Us<strong>in</strong>g the correspondence between bisectors and real geodesics, we have that isometries<br />

act transitively over the <strong>space</strong> of bisectors, s<strong>in</strong>ce they do so over the <strong>space</strong> of real geodesics.<br />

It is known that there are no non-trivial isometries which fix po<strong>in</strong>twise a bisector , s<strong>in</strong>ce<br />

they are not totally geodesic hypersurfaces (if ɛ ≠ 0). Anyway, we can consider the reflection<br />

with respect to a slice S of the bisector. This reflexion fixes po<strong>in</strong>twise the slice S and lies the<br />

bisector <strong>in</strong>variant. Moreover, each of these reflexions is also a reflection with respect to the<br />

sp<strong>in</strong>e σ, thus, it fixes the po<strong>in</strong>ts <strong>in</strong> σ ∩ S.


5.1 Space of (k, p)-planes 83<br />

Bisectors are hypersuperfaces of cohomogenity one. That is, the orbit, for the group of the<br />

isometries fix<strong>in</strong>g a bisector, of a po<strong>in</strong>t (except for the po<strong>in</strong>ts <strong>in</strong> the real sp<strong>in</strong>e, which are of null<br />

measure <strong>in</strong> the bisector) <strong>in</strong> a bisector is a submanifold of codimension 1 <strong>in</strong>side the bisector; a<br />

submanifold of codimension 2 <strong>in</strong> CH n (cf. [GG00]).<br />

Thus, bisectors are not homogeneous hypersurfaces. By def<strong>in</strong>ition a hypersurface is homogeneous<br />

if the orbit of each po<strong>in</strong>t (for the isometry group fix<strong>in</strong>g the hypersurface) is all the<br />

hypersurface.<br />

Let us study the orbit of a po<strong>in</strong>t <strong>in</strong> a bisector.<br />

A geodesic γ(t) ⊂ CH n uniquely determ<strong>in</strong>es a tube T (r) of radius r and a bisector E with<br />

real sp<strong>in</strong>e γ.<br />

We study the relation between these two hypersurfaces, obta<strong>in</strong><strong>in</strong>g the orbit of a po<strong>in</strong>t <strong>in</strong><br />

the bisector <strong>in</strong> terms of the tube conta<strong>in</strong><strong>in</strong>g the po<strong>in</strong>t.<br />

Proposition 5.1.8. Let p ∈ E. Consider r such that p ∈ T (r) ∩ E where T (r) denotes the<br />

tube of radius r along the real sp<strong>in</strong>e γ of E. The orbit of p by the isometries fix<strong>in</strong>g E, is given<br />

by T (r) ∩ E.<br />

Proof. Each po<strong>in</strong>t of the orbit O p of p belongs to T (r) s<strong>in</strong>ce the isometries fix<strong>in</strong>g the bisector<br />

fix the sp<strong>in</strong>e, and they preserve distances. Then, O p ⊂ T (r) ∩ E.<br />

Every po<strong>in</strong>t <strong>in</strong> T (r) ∩ E belongs to the orbit of p. Indeed, if q ∈ T (r) ∩ E then d(q, γ(t)) =<br />

d(p, γ(t)), a necessary condition to be q <strong>in</strong> the orbit of p. The projection of the po<strong>in</strong>ts p, q to<br />

Σ can or cannot be the same po<strong>in</strong>t. Let us prove that <strong>in</strong> both cases there exists an isometry<br />

g such that fixes the bisector and g(p) = q.<br />

Suppose that p and q project at Σ to the same po<strong>in</strong>t x. Then p and q belong to the<br />

same slide of E. Let us def<strong>in</strong>e an isometry g fix<strong>in</strong>g x and the bisector. We denote by v the<br />

tangent vector to the real sp<strong>in</strong>e γ at x. As the isometries fix<strong>in</strong>g the bisector also fix γ, g<br />

satisfies dg(v) = ±v. Moreover, as isometries preserve the holomorphic angle dg(Jv) = ±Jv.<br />

Thus, g fixes the <strong>complex</strong> sp<strong>in</strong>e Σ and its orthogonal complement at x, which is the slide<br />

conta<strong>in</strong><strong>in</strong>g p and q, and is isometric to CH n−1 . Now, <strong>in</strong> CH n−1 there exists an isometry ˜g<br />

such that ˜g(p) = q (s<strong>in</strong>ce CH n−1 is a homogeneous <strong>space</strong>). Therefore, g def<strong>in</strong>ed by g(x) = x,<br />

dg(v) = ±v, dg(Jv) = ±Jv and dg(u) = ˜dg(u), for all u ∈ 〈v, Jv〉 ⊥ , gives an isometry of<br />

CK n (ɛ) fix<strong>in</strong>g E and such that g(p) = q.<br />

Suppose that p and q do not project at Σ to the same po<strong>in</strong>t. Let x = Π Σ p and y = Π Σ q.<br />

Note that x, y ∈ γ s<strong>in</strong>ce p and q are po<strong>in</strong>ts <strong>in</strong> the bisector. Then, there exists a reflection ρ<br />

such that ρ(x) = y and ρ(γ) = γ. Thus, dρ takes the orthogonal <strong>space</strong> of {γ x, ′ Jγ x} ′ to the<br />

orthogonal <strong>space</strong> of {γ y, ′ Jγ y}. ′ Moreover, ˜q = ρ(q) satisfies Π Σ˜q = Π Σ q. If we consider the<br />

po<strong>in</strong>ts ˜q and q, then we are <strong>in</strong> the previous case and we know that there exists an isometry g<br />

such that g(˜q) = q.<br />

From this proposition we have that the subset of bisectors conta<strong>in</strong><strong>in</strong>g a po<strong>in</strong>t is a noncompact<br />

set, <strong>in</strong> the <strong>space</strong> of bisectors. In the next proposition, we prove that the measure of<br />

bisectors meet<strong>in</strong>g a regular doma<strong>in</strong> is <strong>in</strong>f<strong>in</strong>ite.<br />

Remark 5.1.9. Denote by dL the <strong>in</strong>variant density of the <strong>space</strong> of bisectors B and by dL 1<br />

the <strong>in</strong>variant density of the <strong>space</strong> of real geodesics <strong>in</strong> CH n . By the correspondence between<br />

geodesics and bisectors we have<br />

dL = dL 1 .<br />

If Σ denotes the <strong>complex</strong> l<strong>in</strong>e conta<strong>in</strong><strong>in</strong>g a real geodesic γ, then the density of the <strong>space</strong><br />

of real geodesics can be expressed as<br />

dL 1 = dL Σ 1 dΣ,


84 Other Crofton formulas<br />

where dΣ denotes the <strong>in</strong>variant density of <strong>complex</strong> l<strong>in</strong>es and dL Σ 1 the <strong>in</strong>variant density of the<br />

<strong>space</strong> of real geodesics conta<strong>in</strong>ed <strong>in</strong> Σ. Us<strong>in</strong>g polar coord<strong>in</strong>ates ρ, θ <strong>in</strong> Σ we get<br />

dL Σ 1 dΣ = cos 4ɛ (ρ)dρdθdΣ. (5.1)<br />

Proposition 5.1.10. The measure of bisectors meet<strong>in</strong>g a regular doma<strong>in</strong> <strong>in</strong> CK n (ɛ) with ɛ < 0<br />

is <strong>in</strong>f<strong>in</strong>ite.<br />

Proof. We prove that the measure of bisectors <strong>in</strong>tersect<strong>in</strong>g a ball B of radius R <strong>in</strong> CK n (ɛ) is<br />

<strong>in</strong>f<strong>in</strong>ite, i.e.<br />

∫<br />

χ(B ∩ L)dL = +∞.<br />

B<br />

Consider the expression for the density of bisectors <strong>in</strong> (5.1). Denote the volume element<br />

of CK n (ɛ) by dx. Fixed a bisector L by x, then dx can be expressed as dx = dx 1 dx L where<br />

dx L denotes the volume element <strong>in</strong> the bisector and dx 1 the length element <strong>in</strong> the direction<br />

N x orthogonal to the bisector at x.<br />

If N y is the normal vector to the bisector at y = Π Σ (x), then the plane spanned by N y<br />

(which co<strong>in</strong>cides with the normal vector to the real sp<strong>in</strong>e <strong>in</strong>side Σ) and the tangent vector u<br />

to the geodesic jo<strong>in</strong><strong>in</strong>g y and x is a totally real plane and conta<strong>in</strong>s N x .<br />

Thus, the plane exp y (span{N y , u}) is isometric to H 2 (ɛ). If r denotes the distance between<br />

y and x, then dx 1 = cos ɛ (r)dy 1 where dy 1 denotes the length element <strong>in</strong> the direction N y .<br />

In the previous remark, we give an expression for dL 1 . Now, we use it tak<strong>in</strong>g polar<br />

coord<strong>in</strong>ates with respect to y ∈ Σ, so ρ = 0. Then, dy 1 = dρ and<br />

dx L dL 1 = dx L dL Σ 1 dΣ = dx L dθdρdΣ = dθdx L dy 1 dΣ =<br />

1<br />

cos ɛ (r) dθdΣdx.<br />

On the other hand, fixed a regular doma<strong>in</strong> Ω ⊂ CK n (ɛ) it follows, for some constant C > 0,<br />

vol(Ω) < 1 C χ(Ω).<br />

Then,<br />

∫<br />

∫<br />

∫ ∫<br />

χ(B ∩ L)dL > C vol(B ∩ L)dL = C dx L dL<br />

B<br />

∫<br />

= C<br />

B<br />

B<br />

∫<br />

(1.16)<br />

= 2πC<br />

L C 1<br />

∫<br />

∫ 2π<br />

B<br />

∫<br />

= 2πvol(B)<br />

0<br />

∫<br />

B<br />

B∩L<br />

∫<br />

1<br />

cos ɛ (r)<br />

∫B<br />

dθdΣdx = 2πC ∫<br />

G C n,n−1<br />

CH n−1 (ɛ)<br />

L C (n−1)[x]<br />

L C 1<br />

1<br />

cos ɛ (r) dΣdx<br />

cos 2 ɛ(r)<br />

cos ɛ (r) dp n−1dG n,n−1 dx<br />

cos ɛ (r)dx = +∞.<br />

5.2 Variation of the measure of planes meet<strong>in</strong>g a regular doma<strong>in</strong><br />

At Chapter 4 we give an expression for the measure of <strong>complex</strong> r-planes meet<strong>in</strong>g a regular<br />

doma<strong>in</strong> <strong>in</strong> CK n (ɛ). Now, we give a generalization of this result for the <strong>space</strong> of (k, p)-planes<br />

<strong>in</strong> C n , and for the <strong>space</strong> of totally real k-planes <strong>in</strong> CK n (ɛ).<br />

First, we need the expression of the density of the <strong>space</strong> of (k, p)-planes with respect to<br />

the <strong>forms</strong> ω ij def<strong>in</strong>ed at (1.13).


5.2 Variation of the measure of planes meet<strong>in</strong>g a regular doma<strong>in</strong> 85<br />

Lemma 5.2.1.<br />

1. In C n , the <strong>space</strong> L k,p is a homogeneous <strong>space</strong> and<br />

L k,p<br />

∼ = U(n) ⋉ C n /(U(p) × O(k − 2p) × U(n − k + p) ⋉ R n ).<br />

Let {g; g 1 , g 2 , ..., g 2n−1 , g 2n } with g 2i = Jg 2i+1 be a J-mov<strong>in</strong>g frame adapted to a (k, p)-<br />

plane <strong>in</strong> g such that {g 1 , Jg 1 , ..., g p , Jg p , g 2p+1 , ..., g 2k−1 } expand the tangent <strong>space</strong> to the<br />

(k, p)-plane. The <strong>in</strong>variant density of L k,p is given by<br />

∣ ∣∣∣∣∣ ∧ ∧<br />

dL k,p =<br />

ω i ω ji (5.2)<br />

∣<br />

i<br />

where i ∈ {2p + 2, 2p + 4, ..., 2k, 2k + 1, 2k + 2, ..., 2n} and j ∈ {1, 3, ..., 2p − 1, 2p + 1, 2p +<br />

3, ..., 2k − 1}.<br />

2. In CK n (ɛ), ɛ ≠ 0, the <strong>space</strong> of <strong>complex</strong> p-planes L C p and the <strong>space</strong> of totally real k-planes<br />

L R k<br />

are homogeneous <strong>space</strong>s and<br />

j,i<br />

L C p ∼ = U ɛ (n)/(U ɛ (p) × U(n − p)),<br />

where<br />

L R k ∼ = U ɛ (n)/(O ɛ (k) × U(n − k)),<br />

U ɛ (n) =<br />

{ U(1 + n), if ɛ > 0,<br />

U(1, n), if ɛ < 0.<br />

, O ɛ (k) =<br />

{ O(1 + k), if ɛ > 0,<br />

O(1, k), if ɛ < 0.<br />

Moreover, fixed a J-mov<strong>in</strong>g frame as <strong>in</strong> the previous statement, the expression (5.2)<br />

rema<strong>in</strong>s true.<br />

Proof. 1. By Lemma 1.5.1 we have that the isometry group of C n acts transitively over<br />

J-basis. Thus, there exists an isometry that carries a fixed (k, p)-plane to another.<br />

The isotropy group of a (k, p)-plane <strong>in</strong> C n is isomorphic to U(p)×O(k−2p)×U(n−k+p)<br />

s<strong>in</strong>ce (k, p)-planes <strong>in</strong> C n are totally geodesic submanifolds and the tangent <strong>space</strong> at each<br />

po<strong>in</strong>t is isometric to C p ⊕ R k−2p .<br />

The density can be obta<strong>in</strong>ed us<strong>in</strong>g the theory of mov<strong>in</strong>g frames that we have discussed<br />

<strong>in</strong> Section 1.3.<br />

2. The arguments for the previous case are also valid, s<strong>in</strong>ce we restrict to totally geodesic<br />

submanifolds.<br />

The follow<strong>in</strong>g proposition is a generalization of the Proposition 4.2.1 for any (k, p)-plane<br />

<strong>in</strong> C n .<br />

Proposition 5.2.2. Let Ω ⊂ C n be a regular doma<strong>in</strong>, X a smooth vector field def<strong>in</strong>ed at C n<br />

with φ t the flow associated to X and Ω t = φ t (Ω). Then, <strong>in</strong> the <strong>space</strong> of (k, p)-planes L k,p <strong>in</strong><br />

C n , it is satisfied<br />

∫<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L k,p )dL k,p = 〈∂φ/∂t, N〉<br />

σ k (II| V )dV dx<br />

t=0<br />

∫L k,p<br />

∂Ω<br />

G n,k,p (T x∂Ω)<br />

where N is the outward normal field at ∂Ω and σ k (II| V ) denotes the k-th symmetric elementary<br />

function of II restricted to V ∈ G n,k,p (T x ∂Ω), the Grassmanian of the (k, p)-planes conta<strong>in</strong>ed<br />

<strong>in</strong> the tangent <strong>space</strong> of ∂Ω at x.


86 Other Crofton formulas<br />

It is also satisfied the follow<strong>in</strong>g extension of Proposition 4.2.1 for totally real planes <strong>in</strong><br />

CK n (ɛ).<br />

Proposition 5.2.3. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>, X a smooth vector field def<strong>in</strong>ed at<br />

CK n (ɛ) with φ t the flow associated to X and Ω t = φ t (Ω). Then, <strong>in</strong> the <strong>space</strong> of totally real<br />

k-plane L R k <strong>in</strong> CKn (ɛ), it is satisfied<br />

∫<br />

d<br />

dt∣ t=0<br />

L R k<br />

∫<br />

∫<br />

χ(Ω t ∩ L k,p )dL k,p = 〈∂φ/∂t, N〉<br />

σ k (II| V )dV dx<br />

∂Ω<br />

G n,k,0 (T x∂Ω)<br />

where N is the outward normal field at ∂Ω and σ k (II| V ) denotes the k-th symmetric elementary<br />

function of II restricted to V ∈ G n,k,0 (T x ∂Ω), the Grassmanian of the totally real k-planes<br />

conta<strong>in</strong>ed <strong>in</strong> the tangent <strong>space</strong> of ∂Ω at x.<br />

Proof. This proof is analog to the proof of Proposition 4.2.1, s<strong>in</strong>ce the expression for the density<br />

of the <strong>space</strong> of totally real planes <strong>in</strong> (5.2) holds. We just have to modify the construction of<br />

the map γ <strong>in</strong> (4.7).<br />

For every x ∈ ∂Ω consider the curve c(t) = ϕ t (x). For every t, let D c(t) = 〈N c(t) , JN c(t) 〉 ⊥ ⊂<br />

dϕ t (T x ∂Ω) the <strong>complex</strong> hyperplane tangent to ϕ t (∂Ω) at c(t). If ∇ ∂t denotes the covariant<br />

derivative of CK n (ɛ) along c(t), we def<strong>in</strong>e<br />

∇ D ∂t X(t) = π t(∇ ∂t X(t))<br />

where π t : T c(t) CK n (ɛ) → D c(t) denotes the orthogonal projection. Given a vector X ∈ T x ∂Ω,<br />

there exists a unique vector field X(t) def<strong>in</strong>ed along c(t) such that ∇ D ∂tX(t) = 0 (it can be<br />

proved <strong>in</strong> the same way as the existence of the usual parallel translation). This def<strong>in</strong>e a l<strong>in</strong>ear<br />

map ψ t : D x → D c(t) , which preserves the <strong>complex</strong> structure J s<strong>in</strong>ce<br />

∇ D ∂t JX(t) = π t(∇ ∂t JX(t)) = π t (J∇ ∂t X(t)) = Jπ t (∇ ∂t X(t)).<br />

F<strong>in</strong>ally, we extend ψ t l<strong>in</strong>early to ψ t : T x ∂Ω → dϕ t (T x ∂Ω) such that ψ t (JN x ) = JN c(t) . This<br />

map takes totally real planes <strong>in</strong>to totally real planes. So, we can def<strong>in</strong>e the new map γ as<br />

γ : G n,k,p (T ∂Ω) × (−ɛ, ɛ) −→ L k,p<br />

((x, V ), t) ↦→ exp φt(x) ψ t (V ) .<br />

5.3 Measure of real geodesics <strong>in</strong> CK n (ɛ)<br />

The follow<strong>in</strong>g result, obta<strong>in</strong>ed straightforward from the last proposition, states that <strong>in</strong> <strong>complex</strong><br />

<strong>space</strong> <strong>forms</strong>, the measure of real geodesics meet<strong>in</strong>g a regular doma<strong>in</strong> is a multiple of the area<br />

of the doma<strong>in</strong> (as <strong>in</strong> real <strong>space</strong> <strong>forms</strong>).<br />

Theorem 5.3.1. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>, let X be a smooth vector field over<br />

CK n (ɛ), let φ t be the flow associated to X and let Ω t = φ t (Ω). Then<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L 1 )dL R 1 = O 2n+1 ( ˜B 2n−2,n−2 (Ω) + ˜Γ 2n−2,n−1 (Ω))<br />

t=0<br />

L R 1<br />

and<br />

∫<br />

L R 1<br />

χ(Ω ∩ L 1 )dL R 1 = ω 2n µ 2n−1,n−1 (Ω) = ω 2n<br />

2 vol(∂Ω).


5.4 Measure of real hyperplanes <strong>in</strong> C n 87<br />

Proof. From Proposition 5.2.2 we have<br />

∫<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L 1 )dL R 1 =<br />

t=0<br />

L R 1<br />

∂Ω<br />

∫<br />

〈∂φ/∂t, N〉<br />

σ 1 (II| V )dV dx.<br />

G R n,1,0 (Tx∂Ω)<br />

Let us study the <strong>in</strong>tegral with respect to the Grassmanian of geodesics <strong>in</strong> the tangent <strong>space</strong><br />

of each po<strong>in</strong>t x ∈ ∂Ω. We denote by {f 1 , . . . , f 2n−1 } the pr<strong>in</strong>cipal directions at x. Then,<br />

∫<br />

σ 1 (II| V )dV = 1 ∫<br />

k n (v)dv = 1 2n−1<br />

∑<br />

∫<br />

〈v, f i 〉 2 k i dv<br />

G R n,1,0 (Tx∂Ω) 2 S 2n−1 2<br />

i=1<br />

S 2n−1<br />

= 1 2n−1<br />

∑<br />

∫<br />

k i 〈v, f i 〉 2 dv (3.1.2)<br />

= O 2n−1<br />

∑<br />

2n−1<br />

k i = O 2n−1(2n − 1)<br />

tr(II).<br />

2<br />

S 2n−1 4n<br />

4n<br />

i=1<br />

Thus, by Examples 2.4.20.3 and 2.4.20.4<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L 1 )dL R 1 = O ∫<br />

2n−1(2n + 1)<br />

〈X, N〉tr(II)dx<br />

t=0 L R 4n<br />

1<br />

∂Ω<br />

= O ∫<br />

(<br />

2n−1<br />

〈X, N〉<br />

4n<br />

= O 2n−1<br />

4 n!<br />

=<br />

ω 2n<br />

2 (n − 1)!<br />

N(Ω)<br />

i=1<br />

1<br />

(n − 1)! γ ∧ θn−1 2 +<br />

( ∫<br />

〈X, N〉γ ∧ θ2 n−1 + (n − 1)<br />

N(Ω)<br />

N(Ω)<br />

(˜Γ′ 2n−2,n−1 (Ω) + (n − 1) ˜B 2n−2,n−2(Ω))<br />

′<br />

= O 2n+1 ( ˜B 2n−2,n−2 (Ω) + ˜Γ 2n−2,n−1 (Ω)).<br />

)<br />

1<br />

(n − 2)! β ∧ θ 1 ∧ θ2<br />

n−2<br />

∫<br />

)<br />

〈X, N〉β ∧ θ 1 ∧ θ n−2<br />

In C n , the valuation ∫ L<br />

χ(∂Ω∩L R 1 )dL 1 has degree 2n−1, so, it is a multiple of B 2n−1,n−1 (Ω),<br />

1<br />

which has variation (cf. Proposition 4.1.7)<br />

δ X µ 2n−1,n−1 (Ω) = c n,2n−1,n−1 (2c −1<br />

n,2n−2,n−1˜Γ 2n−2,n−1 (Ω) + c −1<br />

n,2n−2,n−2 (n − 1) ˜B 2n−2,n−2 (Ω))<br />

= c n,2n−2,n−1 (˜Γ ′ 2n−2,n−1(Ω) + (n − 1) ˜B 2n−2,n−2(Ω))<br />

′<br />

1<br />

= (˜Γ ′<br />

(n − 1)!ω<br />

2n−2,n−1(Ω) + (n − 1) ˜B 2n−2,n−2(Ω)).<br />

′<br />

1<br />

Therefore, compar<strong>in</strong>g both variations we get the stated result <strong>in</strong> C n . The same expression<br />

holds for ɛ ≠ 0 s<strong>in</strong>ce the variation of µ 2n−1,n−1 does not depend on ɛ.<br />

The relation with the volume of the boundary of the doma<strong>in</strong> is obta<strong>in</strong>ed from the relation<br />

of β 2n−1,n−1 with the second fundamental form given <strong>in</strong> Example 2.4.20.5.<br />

5.4 Measure of real hyperplanes <strong>in</strong> C n<br />

The measure of real hyperplanes <strong>in</strong>tersect<strong>in</strong>g a regular doma<strong>in</strong> <strong>in</strong> C n also follows immediately<br />

from Proposition 5.2.2. This particular case has <strong>in</strong>terest by its own s<strong>in</strong>ce real hyperplanes are<br />

submanifolds of codimension 1.<br />

Theorem 5.4.1. Let Ω ⊂ C n be a regular doma<strong>in</strong>, X a smooth vector field over C n , φ t the<br />

flow associated to X and Ω t = φ t (Ω). Then<br />

d<br />

dt∣ χ(Ω t ∩ L 2n−1,n−1 )dL 2n−1,n−1 = O 2n+1˜Γ0,0 (Ω)<br />

t=0<br />

∫L 2n−1,n−1<br />

and<br />

∫<br />

L 2n−1,n−1<br />

χ(Ω ∩ L 2n−1,n−1 )dL 2n−1,n−1 = ω 2n−1 µ 1,0 (Ω).<br />

2


88 Other Crofton formulas<br />

Proof. From Proposition 4.2.1 we have<br />

∫ ∫<br />

d<br />

dt∣ χ(Ω t ∩ L 2n−1,n−1 )dL 2n−1,n−1 = 〈X, N〉<br />

σ 2n−1 (II| V )dV dx<br />

t=0<br />

∫L 2n−1,n−1 ∂Ω 0 G n,2n−1,n−1 (T<br />

∫<br />

x∂Ω)<br />

= 〈X, N〉σ 2n−1 (II)dx<br />

∫∂Ω<br />

1<br />

= 〈X, N〉<br />

(n − 1)! γ ∧ θn−1 0<br />

N(Ω)<br />

= 2c−1 n,0,0,<br />

(n − 1)! ˜Γ 0,0 (Ω).<br />

In C n , the valuation ∫ L 2n−1,n−1<br />

χ(∂Ω ∩ L 2n−1,n−1 )dL 2n−1,n−1 has degree 1, thus, it is a<br />

multiple of µ 1,0 , which has variation<br />

δ X µ 1,0 = 2c n,1,0 c −1<br />

n,0,0˜Γ 0,0 .<br />

Then, compar<strong>in</strong>g both expressions we obta<strong>in</strong> the result.<br />

Remark 5.4.2. From Example 2.4.20, it follows the equality<br />

µ 1,0 (Ω) = 1 det(II| D ) =<br />

ω 2n−1<br />

∫∂Ω<br />

1 M<br />

ω<br />

2n−2(∂Ω).<br />

D<br />

2n−1<br />

On the other hand, there is just one l<strong>in</strong>early <strong>in</strong>dependent valuation <strong>in</strong> the <strong>space</strong> of cont<strong>in</strong>uous<br />

translation and U(n)-<strong>in</strong>variant valuations <strong>in</strong> C n of degree 1. Thus,<br />

µ 1,0 (Ω) = cM 2n−2 (∂Ω),<br />

and the measure of real hyperplanes <strong>in</strong> C n meet<strong>in</strong>g a regular doma<strong>in</strong> is a multiple of the<br />

so-called “mean width”, as <strong>in</strong> the Euclidean <strong>space</strong>.<br />

5.5 Measure of coisotropic planes <strong>in</strong> C n<br />

A sub<strong>space</strong>s of C n is called coisotropic if its orthogonal is a totally real plane.<br />

Lemma 5.5.1. The (2n − p, n − p)-planes <strong>in</strong> C n are the coisotropic planes.<br />

Proof. If L ∈ L 2n−p,n−p , then L ⊥ has dimensió 2n − (2n − p) = p. The dimension of the<br />

maximal <strong>complex</strong> sub<strong>space</strong> conta<strong>in</strong>ed <strong>in</strong> L ⊥ is n − (n − p) − p = 0. Thus, L ⊥ is a totally real<br />

plane.<br />

Reciprocally, if L ⊥ is a totally real p-plane, then L has dimension 2n − p and the maximal<br />

<strong>complex</strong> sub<strong>space</strong> has dimension n − p.<br />

Lemma 5.5.2. Let S ⊂ C n be a hypersurface and L ∈ L 2n−p,n−p , p ∈ {1, . . . , n}, be a<br />

(2n−p, n−p)-plane tangent <strong>in</strong> S at p. If N denotes a normal vector to S at x, then JN ∈ T x L.<br />

Proof. As L is a (2n−p, n−p)-plane, we can consider, at each po<strong>in</strong>t, a basis of its tangent <strong>space</strong><br />

of the form {e 1 , Je 1 , . . . , e n−p , Je n−p , e n−p+1 , e n−p+2 , . . . , e n }, <strong>in</strong> a way such that Je i ⊥T x L for<br />

i ∈ {n−p+1, . . . , n}. Moreover, we can complete this basis to a basis of T x C n with the vectors<br />

{J en−p+1 , . . . , Je n }.<br />

On the other hand, at x ∈ L ∩ S, is it satisfied T x L ⊂ T x S, thus, N⊥T x L, i.e.<br />

〈N, e i 〉 = 0, ∀ i ∈ {1, . . . , n}, (5.3)<br />

〈N, Je j 〉 = 0,<br />

∀ j ∈ {1, . . . , n − p}.<br />

Now, if JN = ∑ n<br />

i=1 α ie i + ∑ n<br />

i=1 β iJe i , then N = − ∑ n<br />

i=1 α iJe i + ∑ n<br />

i=1 β ie i . Us<strong>in</strong>g (5.3)<br />

we get JN = ∑ n<br />

i=n−p+1 α ie i . Thus, JN ∈ T x L.


5.5 Measure of coisotropic planes <strong>in</strong> C n 89<br />

From this lemma, we can prove the follow<strong>in</strong>g result.<br />

Theorem 5.5.3. Let Ω ⊂ C n be a regular doma<strong>in</strong>, X a smooth vector field def<strong>in</strong>ed <strong>in</strong> C n , φ t<br />

the flow associated to X and Ω t = φ t (Ω). Then,<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L)dL = vol(G ( )<br />

n,2n−p,n−p)ω 2n−p+1<br />

n<br />

(2n − p + 2)<br />

−1· (5.4)<br />

t=0 L 2n−p,n−p<br />

(n − 1)!<br />

p − 1<br />

and<br />

∫<br />

·<br />

⌊ p−1<br />

2 ⌋ ∑<br />

q=max{0,p−n−1}<br />

4 q−p+1 ( ) 2n + 2q − 2p + 1 −1˜Γp−1,q (Ω),<br />

(2n + 2q − 2p + 3) n + q − p + 1<br />

χ(Ω ∩ L)dL = vol(G n,2n−p,n−p)ω 2n−p<br />

L 2n−p,n−p<br />

(n − 1)!<br />

·<br />

⌊ p 2 ⌋ ∑<br />

q=max{0,p−n}<br />

( ) n −1·<br />

p − 1<br />

( ) 2n + 2q − 2p − 1 −1<br />

4 q−p<br />

n + q − p − 1 2n + 2q − 2p + 1 µ p,q(Ω).<br />

Proof. First of all, we prove that for the <strong>space</strong> of coisotropics planes, the variation of the<br />

measure does not have contribution <strong>in</strong> ˜B p,q . From Proposition 5.2.2 we have<br />

∫<br />

∫<br />

d<br />

dt∣ χ(Ω t ∩ L)dL = 〈∂φ/∂t, N〉<br />

σ 2n−p (II| V )dV dx<br />

t=0<br />

∫L 2n−p,n−p<br />

∂Ω<br />

G n,2n−p,n−p (T x∂Ω)<br />

but each V ∈ G n,2n−p,n−p (T x ∂Ω), by the previous lemma, conta<strong>in</strong>s the JN direction (with N<br />

the outward normal vector to ∂Ω at x), so that II| V always conta<strong>in</strong>s the entry correspond<strong>in</strong>g<br />

to the normal curvature of the direction JN. From Lemma 2.4.18 we have that only the<br />

polynomials obta<strong>in</strong>ed from φ ∗ (γ k,q ) conta<strong>in</strong> this entry of the second fundamental form.<br />

∫<br />

In order to f<strong>in</strong>d the constants, we solve a l<strong>in</strong>ear system. First, note that the functional<br />

L 2n−p,n−p<br />

χ(Ω ∩ L)dL is a valuation <strong>in</strong> C n with homogeneous degree p. Thus, it can be<br />

expressed as a l<strong>in</strong>ear comb<strong>in</strong>ation of the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes with the same degree<br />

∫<br />

L 2n−p,n−p<br />

χ(Ω ∩ L)dL =<br />

⌊ p−1<br />

2 ⌋ ∑<br />

q=max{0,p−n}<br />

A p,q µ p,q (Ω) (5.5)<br />

for some A p,q , which we want to determ<strong>in</strong>e.<br />

Tak<strong>in</strong>g the variation <strong>in</strong> both sides, we f<strong>in</strong>d the value of these constants. By Proposition<br />

4.1.7, the variation on the right hand side of (5.5) is<br />

⌊ p 2 ⌋−1 ∑<br />

q=max{0,p−n−1}<br />

(A p,q 2c n,p,q c −1<br />

n,p−1,q (p − 2q)2 (5.6)<br />

− A p,q+1 2c n,p,q+1 c −1<br />

n,p−1,q (n − p + q + 1)(q + 1))˜Γ p−1,q<br />

+ (A p,q+1 2c n,p,q+1 c −1<br />

n,p−1,q (n − p + q + 3/2)(q + 1)<br />

− A p,q 2c n,p,q c −1<br />

n,p−1,q (p − 2q)(p − 2q − 1)) ˜B p−1,q<br />

+ A p,⌊<br />

p<br />

2 ⌋ 2c n,p,⌊<br />

p<br />

2 ⌋ c n,p−1,⌊<br />

p<br />

2 ⌋ (p − 2⌊ p 2 ⌋)2˜Γp−1,⌊ p<br />

2 ⌋ .<br />

Impos<strong>in</strong>g that the variation vanishes on ˜B p−1,q we get some equations, from which we obta<strong>in</strong><br />

the relations<br />

A p,q+1 = n − p + q + 1<br />

(n − p + q + 3/2) A p,q. (5.7)


90 Other Crofton formulas<br />

So, each A p,q+1 is a multiple of A p,max{0,p−n} . To f<strong>in</strong>d this value, we need another equation,<br />

obta<strong>in</strong>ed tak<strong>in</strong>g II| D = λId, λ ∈ R + , and equation the expression (5.6) with the variation of<br />

the Proposition 5.2.2. Then, for each (n, r) we have a compatible l<strong>in</strong>ear system, s<strong>in</strong>ce constant<br />

<strong>in</strong> (5.5) exist. Moreover, by (5.7) they are unique. Do<strong>in</strong>g so, we get, <strong>in</strong> the same way as <strong>in</strong> the<br />

proof of Proposition 4.3.1, the desired result. F<strong>in</strong>ally, we get the variation substitut<strong>in</strong>g the<br />

obta<strong>in</strong>ed values of A p,q at (5.6).<br />

An <strong>in</strong>terest<strong>in</strong>g particular case of the last theorem is the case of Lagrangian planes. This is<br />

the case stated by Alesker [Ale03] as a remarkable case of study. From the previous theorem<br />

we can give explicitly the constant <strong>in</strong> the theorem of Alesker reproduced at 2.3.5, but with<br />

respect to the Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes def<strong>in</strong>ed by Bernig-Fu, and not directly by the bases<br />

def<strong>in</strong>ed by Alesker.<br />

Corollary 5.5.4. Let Ω be a regular doma<strong>in</strong> <strong>in</strong> C n with piecewise smooth boundary. Then,<br />

∫<br />

L R n<br />

χ(Ω ∩ L)dL = vol(G n,n,0)ω n<br />

n!<br />

⌊ n−1<br />

2 ⌋<br />

∑<br />

q=0<br />

where L R n denotes the <strong>space</strong> of Lagrangian planes <strong>in</strong> C n .<br />

( ) 2q − 1 −1<br />

4 q−n<br />

q − 1 2q + 1 µ n,q(Ω).<br />

5.6 Measure of Lagrangian planes <strong>in</strong> CK n (ɛ)<br />

Us<strong>in</strong>g the same techniques as <strong>in</strong> Chapter 4, it can be proved the follow<strong>in</strong>g result.<br />

Theorem 5.6.1. Let Ω ⊂ CK n (ɛ) be a regular doma<strong>in</strong>, X a smooth vector field def<strong>in</strong>ed at<br />

CK n (ɛ), φ t the flow associated to X and Ω t = φ t (Ω). Then,<br />

∫<br />

d<br />

dt∣ t=0<br />

L R n<br />

and<br />

if n is odd<br />

χ(Ω t ∩ L)dL = vol(G n,n,0 )ω n+1<br />

(n + 2)<br />

n!<br />

⌊ n−1<br />

2 ⌋ ∑<br />

q=0<br />

4 q−n+1<br />

2q + 3<br />

( ) 2q + 1 −1˜Γn−1,q (Ω),<br />

q + 1<br />

∫<br />

L R n<br />

χ(Ω ∩ L)dL = vol(G n,n,0)ω n<br />

n!<br />

n−1<br />

2∑<br />

q=0<br />

( ) 2q − 1 −1<br />

4 q−n<br />

q − 1 2q + 1 µ n,q(Ω), (5.8)<br />

and if n is even<br />

∫<br />

χ(Ω ∩ L)dL = vol(G n,n,0)<br />

· (5.9)<br />

L R n!<br />

n<br />

⎛<br />

n<br />

2∑<br />

( )<br />

· ⎝ 2q − 1 −1<br />

4 q−n n<br />

⎞<br />

ω n<br />

2∑<br />

( ) n −1<br />

q − 1 2q + 1 µ n,q(Ω) + ɛ i 2 −n+1 ω n−2i<br />

n<br />

2 + i µ<br />

n + 1 n+2i,<br />

n<br />

2 +i (Ω) ⎠ .<br />

q=0<br />

i=1<br />

Proof. In the same way as <strong>in</strong> the proof of Theorem 4.3.5, it is enough to prove that the<br />

variation <strong>in</strong> both sides co<strong>in</strong>cides.<br />

The variation on the left hand side of (5.9) and (5.8) co<strong>in</strong>cides and is <strong>in</strong>dependent on ɛ.<br />

Thus, it co<strong>in</strong>cides with the variation <strong>in</strong> (5.4).<br />

We compute the variation on the right hand side by us<strong>in</strong>g Proposition 4.1.7. Here, we just<br />

reproduce the computations when n is odd. For n even, a similar, but longer study can be<br />

done to verify expression (5.9).


5.6 Measure of Lagrangian planes <strong>in</strong> CK n (ɛ) 91<br />

Denote by E n (Ω) the right hand side of (5.8). Then, by (5.4) we have<br />

δ X E n (Ω)= vol(G n,n,0)ω n<br />

n!<br />

n−1<br />

2∑<br />

q=0<br />

( ) 2q − 1 −1<br />

4 q−n<br />

q − 1 2q + 1 2c n,n,q·<br />

(<br />

· c −1<br />

n,n−1,q (n − 2q)2˜Γn−1,q − c −1<br />

n,n−1,q−1 q2˜Γn−1,q−1<br />

+ c −1<br />

n,n−1,q−1 q(q + 1/2) ˜B n−1,q−1 − c n,n−1,q (n − 2q)(n − 2q − 1) ˜B n−1,q<br />

+ ɛ(c −1<br />

n,n+1,q+1 (n − 2q)(n − 2q − 1) ˜B n+1,q+1 − c −1<br />

n,n+1,q q(q + 1/2) ˜B n+1,q )<br />

n−3 (<br />

)<br />

= 2 vol(G n,n,0)ω n<br />

2∑ c n,n,q 4 q−n (n − 2q) 2<br />

{ {<br />

n!<br />

(2q + 1) ( )<br />

2q−1<br />

− c n,n,q+14 q−n+1 (q + 1) 2<br />

q=0<br />

q−1<br />

(2q + 3) ( )<br />

2q+1<br />

q<br />

(<br />

)<br />

c n,n,q+1 4 q−n+1 (q + 1)(q + 3/2)<br />

+<br />

(2q + 3) ( )<br />

2q+1<br />

− c n,n,q4 q−n (n − 2q)(n − 2q − 1)<br />

q<br />

(2q + 1) ( )<br />

2q−1<br />

q−1<br />

+ɛ<br />

(<br />

2∑ c n,n,q−1 4 q−n−1 (n − 2q + 2)(n − 2q + 1)<br />

(2q − 1) ( )<br />

2q−3<br />

− c n,n,q4 q−n q(q + 1 2 )<br />

)<br />

q−2<br />

(2q + 1) ( )<br />

2q−1<br />

q−1<br />

n−1<br />

q=1<br />

)<br />

c −1<br />

n,n−1,q ˜Γ n−1,q<br />

c −1<br />

n,n−1,q ˜B n−1,q }<br />

c −1<br />

n,n+1,q ˜B n+1,q }.<br />

In order to prove the result, it suffices to prove that this expression is <strong>in</strong>dependent on ɛ. As<br />

for ɛ = 0 we know that δ X E n (Ω) co<strong>in</strong>cides with (5.4), we get the result.<br />

Now, to prove the <strong>in</strong>dependence of ɛ, we collect the coefficient for each ˜B n−1,q and ˜B n+1,q ,<br />

and we prove that they vanish.


Appendix<br />

This appendix conta<strong>in</strong>s a constructive proof of Theorems 4.3.5 and 4.4.1.<br />

Proof of Theorem 4.3.5<br />

We prove that it is possible to f<strong>in</strong>d constants α k,q such that<br />

∫<br />

L C r<br />

χ(Ω ∩ L r )dL r = ∑ k,q<br />

⌊n/2⌋<br />

∑<br />

α k,q B k,q (Ω) + α 2j,j Γ 2j,j (Ω) + α 2n,n vol(Ω)<br />

j=1<br />

(A.10)<br />

where max{0, k − n} ≤ q < k/2 ≤ n.<br />

For ɛ = 0, the existence of these constants follows from the fact that Hermitian <strong>in</strong>tr<strong>in</strong>sic<br />

volumes constitute a basis of smooth valuations. If ɛ ≠ 0, we cannot ensure this fact. Anyway,<br />

we f<strong>in</strong>d the value of the previous constants impos<strong>in</strong>g that the variation <strong>in</strong> both sides of (A.10)<br />

co<strong>in</strong>cides. This is enough to prove (A.10). Indeed, take a deformation Ω t of Ω such that Ω t<br />

converges to a po<strong>in</strong>t. Then, both sides of (A.10) have the same variation and it vanishes <strong>in</strong><br />

the limit.<br />

The variation of the left hand side of (A.10) is given <strong>in</strong> Corollary 4.3.3. The variation of<br />

the right hand side can be computed us<strong>in</strong>g Proposition 4.1.7 and δ X vol = 2 ˜B 2n−1,n−1 .<br />

In the variation of the left hand side, just appear the terms { ˜B 2n−2r−1,q } q . Thus, the<br />

variation of the right hand side can only have these terms. On the other hand, the variation<br />

of a Hermitian <strong>in</strong>tr<strong>in</strong>sic volume B k,q with k even (resp. odd) has only terms ˜B a,b and ˜Γ a ′ ,b ′<br />

with a, a ′ odd (resp. even) (cf. Proposition 4.1.7). As the variation of the left hand side has<br />

only non-vanish<strong>in</strong>g terms with odd subscript, we just consider the valuations with first even<br />

subscript. Do<strong>in</strong>g also the change <strong>in</strong> (4.12), expression (A.10) reduces to<br />

∫<br />

L C r<br />

n−1<br />

∑<br />

(<br />

k=1<br />

n−1<br />

∑<br />

χ(Ω ∩ L r )dL r = (<br />

k=1<br />

∑k−1<br />

q=max{0,2k−n}<br />

C 2k,q B 2k,q ′ (Ω) + D 2k,kΓ ′ 2k,k (Ω)) + dvol(Ω). (A.11)<br />

Now, we start the study to f<strong>in</strong>d constants C k,q , D 2q,q , d such that<br />

∑k−1<br />

q=max{0,2k−n}<br />

⎛<br />

= vol(GC n−1,r )ω 2r+1(r + 1)<br />

)( ⎝<br />

n<br />

r)<br />

( n−1<br />

r<br />

C 2k,q δB 2k,q ′ (Ω) + D 2k,kδΓ ′ 2k,k (Ω) + dδvol(Ω)<br />

n−r−1<br />

∑<br />

q=max{0,n−2r−1}<br />

⎞<br />

( )<br />

2n − 2r − 2q − 1 1<br />

n − r − q 4 ˜B ′ n−r−q−1 2n−2r−1,q(Ω) ⎠ .<br />

By Proposition 4.1.7, this equation gives rise to a l<strong>in</strong>ear system. We write this l<strong>in</strong>ear system<br />

<strong>in</strong> matrix form Ax = b. Consider the vector of unknowns as<br />

x t = (C 2,0 , D 2,1 , C 4,0 , C 4,1 , D 4,2 , . . . , C 2c,max{0,2c−n} , . . . , D 2c,c , . . . , C 2n−2,n−2 , D 2n−2,n−1 , d).<br />

93


94 Appendix<br />

Vector b conta<strong>in</strong>s the coefficient of ˜B ′ k,q , ˜Γ ′ k,q<br />

given <strong>in</strong> (4.23), that is<br />

b t = vol(GC n−1,r ) ( ( )( ) ( )( n − r r + 1 n − r r + 1<br />

n! ( )<br />

n−1<br />

0, . . . , 0,<br />

,<br />

1 1 2 2<br />

r<br />

) 1<br />

2 , . . . , ( n − r<br />

r + 1<br />

)( r + 1<br />

r + 1<br />

) )<br />

r + 1<br />

4 r , 0, . . . , 0 .<br />

Note that b has all entries null except the ones correspond<strong>in</strong>g to ˜B 2n−2r−1,q ′ .<br />

The coefficients of the matrix A conta<strong>in</strong> the variation of each B<br />

2k,q ′ and Γ′ 2q,q with respect<br />

to ˜B r,s ′ and ˜Γ ′ r,s. We denote by (δB<br />

k,q ′ , ˜B r,s ), the coefficient of ˜Br,s <strong>in</strong> the variation of the<br />

valuation B<br />

k,q ′ . By Proposition 4.1.7<br />

(δB<br />

k,q ′ , ˜B r,s) ′ =<br />

(δB<br />

k,q ′ , ˜Γ ′ r,s) =<br />

(δΓ ′ 2q,q , ˜B r,s) ′ =<br />

(δΓ ′ 2q,q , ˜Γ ′ r,s) =<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

⎧<br />

⎨<br />

⎩<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

⎧<br />

⎨<br />

⎩<br />

2q(n + q − k + 1/2), if r = k − 1, s = q − 1<br />

−2(k − 2q)(k − 2q − 1), if r = k − 1, s = q<br />

2ɛ(k − 2q)(k − 2q − 1), if r = k + 1, s = q + 1<br />

−2ɛ(n − k + q)(q + 1/2), if r = k + 1, s = q<br />

0, otherwise.<br />

(k − 2q) 2 , if r = k − 1, s = q<br />

−(n + q − k)q, if r = k − 1, s = q − 1<br />

0, otherwise.<br />

4q(n − q + 1/2), if r = 2q − 1, s = q − 1<br />

−4ɛ((n − q)(2q + 3/2) − (q + 1)/2), if r = 2q + 1, s = q<br />

4ɛ 2 (n − q − 1)(q + 3/2), if r = 2q + 3, s = q + 1<br />

0, otherwise.<br />

−2(n − q)q, if r = 2q − 1, s = q − 1<br />

2ɛ(n − q − 1)(q + 1), if r = 2q + 1, s = q<br />

0, otherwise.<br />

Each column of the matrix A conta<strong>in</strong>s the variation of a valuation B<br />

2k,q ′ , Γ′ 2q,q or the volume.<br />

We take the valuations B<br />

2k,q ′ , Γ′ 2q,q <strong>in</strong> the same order as <strong>in</strong> the vector b. (The volume corresponds<br />

to the last column.) That is, the columns of A conta<strong>in</strong> the variation of the valuations<br />

<strong>in</strong> the follow<strong>in</strong>g order<br />

(δB ′ 2,0, δΓ ′ 2,1, δB ′ 4,0, δB ′ 4,1, δΓ ′ 4,2, . . . , δB ′ 2n−2,n−2, δΓ ′ 2n−2,n−1, δvol).<br />

We denote<br />

δB ′ 2k,· = (δB′ 2k,max{0,2k−n} , δB′ 2k,max{0,2k−n}+1 , . . . , δB′ 2k,k−1 ),<br />

δΓ ′ 2k,· = δΓ′ 2k,k ,<br />

˜B ′ 2k+1,· = ( ˜B ′ 2k+1,max 0,2k−n+1 , ˜B ′ 2k+1,max 0,2k−n+1+1 , . . . , ˜B ′ 2k+1,k )t ,<br />

˜Γ ′ 2k+1,· = (˜Γ ′ 2k+1,max 0,2k−n+1 , ˜Γ ′ 2k+1,max 0,2k−n+1+1 , . . . , ˜Γ ′ 2k+1,k )t .


Proof of Theorem 4.3.5 95<br />

Then, A has the follow<strong>in</strong>g boxes structure<br />

δB 2,· ′ δΓ ′ 2,· δB 4,· ′ δΓ ′ 4,· δB 6,· ′ δΓ ′ 6,· δB 8,· ′ δΓ ′ 8,· · · · δB 2n−4,· ′ δΓ ′ 2n−4,· δB 2n−2,· ′ δΓ ′ 2n−2,· δvol<br />

˜B 1,· ′ ∗ ∗<br />

˜Γ ′ 1,· ∗ ∗<br />

˜B 3,· ′ ∗ ɛ ∗ ɛ ∗ ∗<br />

˜Γ ′ 3,· ∗ ɛ ∗ ∗<br />

˜B 5,· ′ ∗ ɛ ∗ ɛ ∗ ɛ ∗ ∗<br />

˜Γ ′ 5,· ∗ ɛ ∗ ∗<br />

˜B 7,· ′ ∗ ɛ ∗ ɛ ∗ ɛ ∗ ∗<br />

˜Γ ′ 7,· ∗ ɛ ∗ ∗<br />

˜B 9,· ′ ∗ ɛ ∗ ɛ ∗ ɛ<br />

˜Γ ′ 9,·<br />

∗ ɛ<br />

.<br />

˜B 2n−3,· ′ ∗ ɛ ∗ ɛ ∗ ∗<br />

˜Γ ′ 2n−3,· ∗ ɛ ∗ ∗<br />

˜B 2n−1,· ′ ∗ ɛ ∗ ɛ ∗ ɛ ∗<br />

We denoted by ∗ the boxes of A with non-null coefficients and <strong>in</strong>dependent of ɛ, and by ∗ ɛ the<br />

boxes of A with non-null coefficients (for ɛ ≠ 0) and multiples of ɛ.<br />

The structure by boxes of the l<strong>in</strong>ear system given by A suggests the method of resolution:<br />

we start with the top box, and we get the value of variables C 2,q and D 2,1 . Then we solve the<br />

next bloc with rows ˜B ′ 3,·, ˜Γ ′ 3,·, us<strong>in</strong>g the value of variables C 2,q and D 2,1 . We can cont<strong>in</strong>ue this<br />

process, so that, once we know the value of variables C 2k,q and D 2k,k , we substitute it on the<br />

equations given by the rows ˜B ′ 2k+1,·, ˜Γ ′ 2k+1,·.<br />

Recall that the <strong>in</strong>dependent vector b has all terms null except the ones correspond<strong>in</strong>g to<br />

˜B ′ 2n−2r−1,q . Thus, the l<strong>in</strong>ear system is homogeneous for the first equations until ˜B′ 2n−2r−2,q ,<br />

and we can take C k,q = D 2q,q = 0 whenever k ≤ 2n − 2r − 1.<br />

By Theorem 4.3.1 we have a solution for the system Ax = b when ɛ = 0. This solution<br />

has C 2n−2r,q and D 2n−2r,n−r as non-null terms, and satisfies the equations until rows<br />

˜B 2n−2r,·, ′ ˜Γ ′ 2n−2r,· also for ɛ ≠ 0.<br />

So, we consider for all ɛ ∈ R and for all a ∈ {1, ..., m<strong>in</strong>{n − r, r}}<br />

C 2n−2r,n−r−a = vol(GC n−1,r )<br />

4 a n!<br />

D 2n−2r,n−r = vol(GC n−1,r )<br />

2 n!<br />

( ) n − 1 −1 ( n − r<br />

r a<br />

( ) n − 1 −1<br />

.<br />

r<br />

)( r<br />

a)<br />

,<br />

Now, we go on with the resolution of the l<strong>in</strong>ear system <strong>in</strong> CK n (ɛ). We study for each<br />

c ∈ {n − r + 1, . . . , n} the submatrix of A with all rows ˜B 2c−1,q ′ , ˜Γ ′ 2c−1,q . This matrix has the<br />

follow<strong>in</strong>g non-zero columns of A: δΓ ′ 2c−4,c−2 , δB′ 2c−2,·, δΓ′ 2c−2,c−1 , δB′ 2c,· and δΓ′ 2c,c .<br />

Suppose that we know the value of D 2c−4,c−2 , C 2c−2,q and D 2c−2,c−1 . Then, we can substitute<br />

them <strong>in</strong> the equations given by the rows correspond<strong>in</strong>g to ˜B 2c−1,q , ˜Γ 2c−1,q . We get<br />

equations with C 2c,q and D 2c,c as unknowns. If we denote i = max{0, 2c − n}, the matrix of<br />

the coefficients for the obta<strong>in</strong>ed equations (which corresponds to a matrix bloc of A <strong>in</strong>depen-


96 Appendix<br />

dent of ɛ) is<br />

δB 2c,i δB 2c,i+1 δB 2c,i+2 · · · δB 2c,c−2 δB 2c,c−1 δΓ 2c,c<br />

˜B 2c−1,i −4c(2c − 1) 2(n − 2c + 3/2)<br />

˜B 2c−1,i+1 −2(2c − 2)(2c − 3) 4(n − 2c + 5/2)<br />

˜B 2c−1,i+2 −2(2c − 4)(2c − 5)<br />

.<br />

˜B 2c−1,c−2 −24 2(c − 1)(n − c − 1 ) 2<br />

˜B 2c−1,c−1 −4<br />

˜Γ 2c−1,0 (2c) 2 −(n − 2c + 1)<br />

4c(n − c + 1 ) 2<br />

˜Γ 2c−1,1 (2c − 2) 2 −2(n − 2c + 2)<br />

˜Γ 2c−1,2 (2c − 4) 2<br />

.<br />

˜Γ 2c−1,c−2 4 2 −(c − 1)(n − 1)<br />

˜Γ 2c−1,c−1 4 −2c(n − c)<br />

The <strong>in</strong>dependent term is obta<strong>in</strong>ed from the <strong>in</strong>itial <strong>in</strong>dependent term b (which <strong>in</strong> these<br />

cases is always zero) and from the part of the <strong>in</strong>itial equation <strong>in</strong> which we substituted the<br />

value of D 2c−4,c−2 , C 2c−2,q , D 2c−2,c−1 . Compar<strong>in</strong>g the box structure of A on page 95 and its<br />

coefficients on page 94, we obta<strong>in</strong> that the <strong>in</strong>dependent term of this new l<strong>in</strong>ear system has<br />

zero the terms ˜Γ ′ 2c−1,q with q ∈ {max{0, 2c − n}, ..., c − 2}, and the term ˜Γ 2c−1,c−1 equals to<br />

2ɛ(n − c)cD 2c−2,c−1 .<br />

Now, we consider the equations given by rows ˜Γ ′ 2c−1,q and ˜B 2c−1,max{0,2c−n} <strong>in</strong> the previous<br />

matrix, which give a compatible l<strong>in</strong>ear system with one solution. The <strong>in</strong>dependent terms of<br />

the equation given by equation ˜B 2c−1,max{0,2c−n} is ɛ(n−2c+2)C 2c−2,max{0,2c−n−2} . The other<br />

ones are zero. Solv<strong>in</strong>g this system we get the variables C 2c,max{0,2c−n} and D 2c,c <strong>in</strong> terms of<br />

C 2c−2,max{0,2c−n−2} and D 2c−2,c−1 , which we suppose known.<br />

In order to avoid consider<strong>in</strong>g the maximum max{0, 2c − n − 2} we dist<strong>in</strong>guish two cases.<br />

First stage: 2c ≤ n. This case appears if 2r > n (s<strong>in</strong>ce c ∈ {n − r + 1, . . . , n}).<br />

The l<strong>in</strong>ear system we have to solve is given by the augmented matrix<br />

⎛<br />

⎞<br />

−4c(2c − 1) 2(n − 2c + 3/2) ɛ(n − 2c + 2)C 2c−2,0<br />

(2c) 2 −(n − 2c + 1)<br />

(2c − 2) 2 −2(n − 2c + 2)<br />

.<br />

⎜<br />

.. ⎟<br />

⎝<br />

⎠<br />

−(c − 1)(n − c − 1)<br />

4 −2c(n − c) −2ɛc(c − n)D 2c−2,c−1<br />

with variables {C 2c,0 , C 2c,1 , . . . , C 2c,c−1 , D 2c,c }.<br />

From the first two equations we obta<strong>in</strong><br />

ɛ(n − 2c + 2)(n − 2c + 1)<br />

C 2c,0 = C 2c−2,0 ,<br />

4c(n − c + 1)<br />

ɛ(n − 2c + 2)c<br />

C 2c,1 = C 2c−2,0 .<br />

(n − c + 1)<br />

For every q ∈ {0, ..., c − 2} the follow<strong>in</strong>g relations are satisfied<br />

and we get<br />

C 2c,q+1 =<br />

(2c − 2q) 2 C 2c,q = (q + 1)(n − 2c + q + 1)C 2c,q+1 ,<br />

4C 2c,c−1 − 2c(n − c)D 2c,c = −2ɛc(n − c)D 2c−2,c−1 ,<br />

ɛ4 q (n − 2c + 2)!c!(c − 1)!<br />

(n − c + 1)(q + 1)!(n − 2c + q + 1)!(c − q − 1)!(c − q − 1)! C 2c−2,0,<br />

D 2c,c = ɛ(D 2c−2,c−1 +<br />

2(n − 2c + 2)!(c − 1)!4c−2<br />

C 2c−2,0 ).<br />

(n − c + 1)!<br />

(A.12)


Proof of Theorem 4.3.5 97<br />

As the known constants are C 2n−2r,q and D 2n−2r,n−r we write the previous ones <strong>in</strong> terms<br />

of these, us<strong>in</strong>g the recurrence we just obta<strong>in</strong>ed.<br />

C 2c,0 =<br />

ɛ(n − 2c + 2)(n − 2c + 1)<br />

C 2c−2,0<br />

4c(n − c + 1)<br />

= ɛc−(n−r) (n − 2c + 2)(n − 2c + 1) · ... · (n − 2n + 2r − 2 + 2)(n − 2n + 2r − 2 + 1)<br />

4 c−(n−r) c(c − 1) · ... · (n − r + 1)(n − c + 1)(n − c + 2) · ... · (n − (n − r))<br />

(A.13)<br />

= ɛc−(n−r) (2r − n)!(n − r)!(n − c)!vol(G C n−1,r ) ( ) n − 1 −1 ( ) r<br />

4 c−(n−r) (n − 2c)!c!r!4 n−r n!<br />

r n − r<br />

= ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1 ( ) n − c<br />

4 c ,<br />

n!<br />

r c<br />

C 2c,q+1 =<br />

C 2n−2r,0<br />

ɛ(n − 2c + 2)4 q (n − 2c + 1)!c!(c − 1)!<br />

(n − c + 1)(q + 1)!(n − 2c + q + 1)!(c − q − 1)!(c − q − 1)! C 2c−2,0 (A.14)<br />

= ɛ(n − 2c + 2)4q (n − 2c + 1)!c!(c − 1)!ɛ c−1−(n−r) vol(G C n−1,r )(n − c + 1)!<br />

(n − c + 1)(q + 1)!(n − 2c + q + 1)!((c − q − 1)!) 2 4 c−1 n!(c − 1)!(n − 2c + 2)!<br />

ɛ c−(n−r) c!(n − c)!vol(G C n−1,r<br />

=<br />

)<br />

( ) n − 1 −1<br />

4 c−q−1 (q + 1)!(n − 2c + q + 1)!(c − q − 1)!(c − q − 1)!n! r<br />

= ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1 ( )( )<br />

c n − c<br />

4 c−q−1 ,<br />

n! r q + 1 c − q − 1<br />

( n − 1<br />

)<br />

2(n − 2c + 2)!(c − 1)!4c−2<br />

D 2c,c = ɛ<br />

(D 2c−2,c−1 + C 2c−2,0<br />

(n − c + 1)!<br />

(<br />

2(n − 2c + 2)!(c − 1)!4c−2 ɛ c−1−(n−r) vol(G C ( ) )<br />

n−1,r )(n − c + 1)! n − 1 −1<br />

= ɛ D 2c−2,c−1 +<br />

(n − c + 1)! 4 c−1 n!(c − 1)!(n − 2c + 2)! r<br />

= ɛD 2c−2,c−1 + ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1<br />

2 n!<br />

r<br />

= ɛ c−(n−r) D 2n−2r,n−r + (c − (n − r)) ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1<br />

(A.15)<br />

2 n!<br />

r<br />

= ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1<br />

+ (c − (n − r)) ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1<br />

2 n!<br />

r<br />

2 n!<br />

r<br />

= ɛc−(n−r) vol(G C n−1,r )<br />

( ) n − 1 −1<br />

(c + r − n + 1) .<br />

2 n!<br />

r<br />

Thus, we get the value of the unknowns (<strong>in</strong> the vector x) until position D 2⌊n/2⌋,⌊n/2⌋ .<br />

Second stage: 2c ≥ n. Note that B 2c,q ′ is def<strong>in</strong>ed if q ≥ 2c − n > 0. In this case, 2c ≥ n,<br />

and the system we have to solve has the same structure as <strong>in</strong> the previous case (2c ≤ n) but<br />

with less equations and unknowns. Tak<strong>in</strong>g the same equations as <strong>in</strong> the previous case 2c ≤ n,<br />

we obta<strong>in</strong>, as augmented matrix,<br />

⎛<br />

(2c − n) ɛ((4c − 2n − 1)C 2c−2,2c−n−1−<br />

−4(n − c + 1)(2n − 2c + 1)C 2c−2,2c−n−2)<br />

⎜ (2n − 2c)<br />

⎝<br />

2 −(2c − n − 1)<br />

⎟<br />

⎠ .<br />

4 −2c(n − c) −2ɛc(c − n)D 2c−2,c−1<br />

⎞<br />

r<br />

) −1


98 Appendix<br />

From the first equation, it follows<br />

ɛ<br />

C 2c,2c−n =<br />

2c − n ((4c−2n−1)C 2c−2,2c−n−1 −2(2n−2c+2)(2n−2c+1)C 2c−2,2c−n−2 ). (A.16)<br />

For a ∈ {0, ..., n − c}, by relation<br />

we get<br />

and from<br />

C 2c,2c−n+a =<br />

and (A.17) we get<br />

(2n − 2c − 2a + 2) 2 C 2c,2c−n+a−1 = a(2c − n + a)C 2c,2c−n+a<br />

=<br />

=<br />

4(n − c − a + 1)2<br />

C 2c,2c−n+a−1<br />

a(2c − n + a)<br />

4 a (n − c − a + 1) 2 (n − c − a + 2) 2 . . . (n − c) 2<br />

a(a − 1) . . . 2(2c − n + a)(2c − n + a − 1) . . . (2c − n + 1) C 2c,2c−n<br />

4 a (n − c)!(n − c)!(2c − n)!<br />

a!(n − c − a)!(n − c − a)!(2c − n + a)! C 2c,2c−n<br />

4C 2c,c−1 − 2(n − c)cD 2c,c = −2ɛ(n − c)(c)D 2c−2,c−1<br />

D 2c,c = ɛD 2c−2,c−1 +<br />

2<br />

c(n − c) C 2c,c−1<br />

= ɛD 2c−2,c−1 + 2 · 4n−c−1 (n − c)!(n − c)!(2c − n)!<br />

C 2c,2c−n<br />

c(n − c)(n − c − 1)!(c − 1)!<br />

= ɛD 2c−2,c−1 + 2 4n−c−1 (n − c)!(2c − n)!<br />

C 2c,2c−n .<br />

c!<br />

(A.17)<br />

(A.18)<br />

In order to obta<strong>in</strong> the value of C 2c,2c−n we use the value of C 2c−2,2c−n−1 and C 2c−2,2c−n−2<br />

if c ∈ {n − r, ..., ⌊n/2⌋}. We consider c 0 = ⌊ n+2<br />

2 ⌋.<br />

From the previous case 2c ≤ n we know the value of the unknowns C 2c0 −2,2c 0 −n−1,<br />

C 2c0 −2,2c 0 −n−2 and D 2c0 −2,c 0 −1. For n even we have (we omit the analogous computation<br />

for n odd)<br />

C 2c0 −2,2c 0 −n−1 = C n,1 = ɛn/2−(n−r) vol(G C n−1,r )(n/2)!(n/2)! ( ) n − 1 −1<br />

4 n/2−1 (A.19)<br />

n!((n − 2)/2)!((n − 2)/2)! r<br />

= ɛr−n/2 vol(G C n−1,r ) ( ) n − 1 −1<br />

2 n n 2 ,<br />

n!<br />

r<br />

C 2c0 −2,2c 0 −n−2 = C n,0 = ɛr−n/2 vol(G C n−1,r )<br />

2 n n!<br />

( ) n − 1 −1<br />

,<br />

r<br />

D 2c0 −2,c 0 −1 = D n,n/2 = ɛn/2−(n−r) vol(G C n−1,r )<br />

(r − n ( ) n − 1 −1<br />

2 n!<br />

2 + 1) .<br />

r<br />

Then<br />

ɛ<br />

C n+2,2 =<br />

2c − n ((4c − 2n − 1)C 2c−2,2c−n−1 − 2(2n − 2c + 2)(2n − 2c + 1)C 2c−2,2c−n−2 )<br />

= ɛr−n/2+1 vol(G C n−1,r )<br />

( ) n − 1 −1<br />

2 n+1 (3n 2 − 2n(n − 1))<br />

n!<br />

r<br />

= ɛr−n/2+1 vol(G C n−1,r ) ( ) n − 1 −1<br />

2 n+1 n(n + 2) .<br />

n!<br />

r


Proof of Theorem 4.3.5 99<br />

Once we know the expression of C n+2,2 we f<strong>in</strong>d the value of C 2c,2c−n , for every c ∈ {⌊n/2⌋, . . . , n},<br />

us<strong>in</strong>g the recurrence for C 2c,2c−n and C 2c−2,2c−n−1 . First, we have<br />

(A.16)<br />

C 2c,2c−n =<br />

(A.17)<br />

ɛ<br />

2c − n ((4c − 2n − 1)C 2c−2,2c−n−1 − 2(2n − 2c + 2)(2n − 2c + 1)C 2c−2,2c−n−2 )<br />

= ɛC 2c−2,2c−n−2<br />

·<br />

(<br />

2c − n<br />

(4c − 2n − 1)4(n − c + 1)!(n − c + 1)!(2c − n − 2)!<br />

·<br />

(n − c)!(n − c)!(2c − n − 1)!<br />

(<br />

4ɛ(n − c + 1) (4c − 2n − 1)(n − c + 1) − (2n − 2c + 1)(2c − n − 1)<br />

=<br />

2c − n<br />

(2c − n − 1)<br />

4ɛ(n − c + 1)c<br />

=<br />

(2c − n)(2c − n − 1) C 2c−2,2c−n−2.<br />

)<br />

− 4(n − c + 1)(2n − 2c + 1)<br />

)<br />

C 2c−2,2c−n−2<br />

We go on with this recurrence until C ∗,∗−n with ∗ ≤ n+2<br />

2<br />

. In this case we know the value of<br />

the constants, and we can f<strong>in</strong>d the value of C 2c,2c−n .<br />

C 2c,2c−n = (4ɛ)c−(n+2)/2 c(c − 1) · ... · ((n + 4)/2)(n − c + 1)(n − c + 2) · ... · (n/2 − 1)<br />

C n+2,2<br />

(2c − n)(2c − n − 1) · ... · 4 · 3<br />

F<strong>in</strong>ally,<br />

= (4ɛ)c−(n+2)/2 c!((n − 2)/2)!2<br />

((n + 2)/2)!(n − c)!(2c − n)! C n+2,2<br />

= (4ɛ)c−(n+2)/2 2 3 ( ) c ɛ r−n/2+1 vol(G C n−1,r ) ( ) n − 1 −1<br />

(n + 2)n 2c − n 2 n+1 n(n + 2)<br />

n!<br />

r<br />

= ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1 ( ) c<br />

4 n−c . (A.20)<br />

n! r 2c − n<br />

and<br />

4 a (n − c)!(n − c)!(2c − n)! ɛ c−(n−r) c!vol(G C n−1,r<br />

C 2c,2c−n+a =<br />

) ( n − 1<br />

a!(n − c − a)!(n − c − a)!(2c − n + a)! 4 n−c (2c − n)!(n − c)!n! r<br />

= ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1 ( )( )<br />

n − c c<br />

4 n−c−a n! r n − c − a 2c − n + a<br />

) −1<br />

= ɛD 2c−2,c−1 + 2 4n−c−1 (n − c)!(2c − n)! ɛ c−(n−r) c!vol(G C n−1,r ) ( n − 1<br />

c! 4 n−c (2c − n)!(n − c)!n! r<br />

= ɛD 2c−2,c−1 + ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1<br />

2 n!<br />

r<br />

= ɛ c−n/2 D n,n/2 + (c − n/2) ɛc−(n−r) vol(G C n−1,r ) ( ) n − 1 −1<br />

2 n!<br />

r<br />

= ɛc−(n−r) vol(G C n−1,r )<br />

( ) n − 1 −1<br />

(c + r − n + 1) .<br />

2 n!<br />

r<br />

D 2c,c<br />

(A.17) and (A.20)<br />

) −1<br />

To determ<strong>in</strong>e the value of d, the coefficient of δvol, we consider the last equation of the


100 Appendix<br />

<strong>in</strong>itial l<strong>in</strong>ear system<br />

So,<br />

d<br />

(n − 1)! = −2ɛ2 (2n − 1)D 2n−4,n−2 − 4ɛC 2n−2,n−2 + 2ɛ(3n − 1)D 2n−2,n−1<br />

= 2ɛ r (−(2n − 1)(r − 1) − (n − 1) + (3n − 1)r) vol(GC n−1,r ) ( n − 1<br />

2 n! r<br />

= ɛr vol(G C n−1,r ) ( ) n − 1 −1<br />

n(r + 1) .<br />

n!<br />

r<br />

( ) n − 1 −1<br />

d = ɛ r vol(G C n−1,r)(r + 1) .<br />

r<br />

Now, we have to prove that the given solution satisfies all the equations we did not use to<br />

solve the system. This is because we cannot ensure that the equation (A.10) has solution.<br />

Let us study first the case 2c ≤ n. Consider the matrix on page 96. The rows we did not<br />

use correspond to ˜B 2c−1,q ′ , q ∈ {1, ..., c − 1}. Suppose q ≠ c − 1. The equations given by this<br />

row are ˜B 2c−1,q<br />

′<br />

−(2c − 2q)(2c − 2q − q)C 2c,q + (q + 1)(n − 2c + q + 3/2)C 2c,q+1<br />

= −ɛ(2c − 2q)(2c − 2q − 1)C 2c−2,q−1 + ɛ(n − 2c + q + 2)(q + 1/2)C 2c−2,q .<br />

For q = c − 1, the equation is<br />

2ɛ 2 (n − c + 1)(c − 1/2)D 2c−4,c−2 + 2ɛC 2c−2,c−2 − 2ɛ((n − c + 1)(2c − 1/2) − c/2)D 2c−2,c−1<br />

= −2C 2c,c−1 + 2c(n − 2 + 1/2)D 2c,c .<br />

Substitut<strong>in</strong>g the value of each C ∗,· and D ∗,· given on page 97 we prove that the equations are<br />

satisfied.<br />

In the same way, we can prove that all equations appear<strong>in</strong>g <strong>in</strong> the case 2c ≥ n are also<br />

satisfied.<br />

F<strong>in</strong>ally, us<strong>in</strong>g aga<strong>in</strong> the relation <strong>in</strong> (4.12), we get the result with respect to {B k,q , Γ k,q }.<br />

Proof of Theorem 4.4.1<br />

The idea of the proof of this theorem is the same as for Theorem 4.3.5. In the same way, if the<br />

variation δ X is the same <strong>in</strong> both sides, for all differentiable vector field X, then the expression<br />

holds.<br />

From the Gauss-Bonnet-Chern formula, we know that χ(Ω) can be written as the <strong>in</strong>tegral<br />

over N(Ω) of a differential form O(2n)-<strong>in</strong>variant, and also U(n)-<strong>in</strong>variant. Thus, by Proposition<br />

2.4.5 there exist constants C k,q , D k,q , d such that<br />

) −1<br />

χ(Ω) = ∑ k,q<br />

⌊ n 2 ⌋<br />

C k,q B k,q ′ (Ω) + ∑<br />

D 2j,j Γ ′ 2j,j(Ω) + dvol(Ω)<br />

j=1<br />

(A.21)<br />

where max{0, k − n} ≤ q < k/2 ≤ n, and B<br />

k,q ′ and Γ′ k,q<br />

are the valuations def<strong>in</strong>ed <strong>in</strong> (4.12).<br />

Tak<strong>in</strong>g the variation <strong>in</strong> both sides of the previous equality we have<br />

0 = ∑ (c k,q ˜B′ k,q (Ω) + d 2q,q ˜Γ′ 2q,q (Ω))<br />

k,q<br />

with c k,q and d k,q l<strong>in</strong>ear comb<strong>in</strong>ation of C k,q and D 2q,q .


Proof of Theorem 4.4.1 101<br />

Thus, we have to impose c k,q = d k,q = 0.<br />

The variation of Γ ′ 0,0 <strong>in</strong> CKn (ɛ) is (cf. Corollary 4.1.9)<br />

δΓ ′ 0,0(Ω) = 2ɛ(−(3n − 1) ˜B ′ 1,0(Ω) + (n − 1)˜Γ ′ 1,0(Ω) + 3ɛ(n − 1) ˜B ′ 3,1(Ω)).<br />

It is necessary to cancel the variation of the terms ˜B 1,0 ′ , ˜B 3,1 ′ and ˜Γ ′ 1,0 . By Proposition 4.1.7 we<br />

have that the variation of a valuation B<br />

k,q ′ <strong>in</strong> CKn (ɛ) with k even (resp. odd) has only terms<br />

˜B<br />

k ′ ′ ,q<br />

and ˜Γ ′ ′ k ′ ,q<br />

with k ′ odd (resp. even). Thus, <strong>in</strong> the expresssion (A.21) we can restrict the<br />

′<br />

value of k to k even, and (A.21) can be reduced to<br />

⎛<br />

n−1<br />

∑<br />

χ(Ω) = ⎝<br />

k=0<br />

∑k−1<br />

q=max{0,2k−n}<br />

C 2k,q B ′ 2k,q (Ω) + D 2k,kΓ ′ 2k,k (Ω) ⎞<br />

⎠ + dvol(Ω).<br />

(A.22)<br />

The right hand side <strong>in</strong> the previous equality co<strong>in</strong>cides with the right hand side of (A.11) plus<br />

the term D 0,0 Γ ′ 0,0 . Thus, the variation is very similar and the l<strong>in</strong>ear system we have to solve<br />

will be also very similar to the one solved <strong>in</strong> Theorem 4.3.5. The only different equations are<br />

the ones given by c 1,0 = 0, d 1,0 = 0 and c 3,1 = 0, that is<br />

−ɛ(3n − 1)D 0,0 − 2C 2,0 + 2(n − 1/2)D 2,1 = 0,<br />

ɛ(n − 1)D 0,0 + 2C 2,0 − (n − 1)D 2,1 = 0,<br />

3ɛ 2 (n − 1)D 0,0 + 2ɛC 2,0 − ɛ(7n − 9)D 2,1 − 2C 4,1 + 2(2n − 3)D 4,2 = 0.<br />

(A.23)<br />

We f<strong>in</strong>d the value of D 0,0 for ɛ = 0, i.e. <strong>in</strong> C n , us<strong>in</strong>g the Gauss-Bonnet formula so that<br />

D 0,0 =<br />

1<br />

O 2n−1 (n − 1)! = 1<br />

2nω 2n (n − 1)! = n!<br />

2 n!π n = 1<br />

2π n .<br />

The choice for the value of D 0,0 ensures that both sides <strong>in</strong> (A.21) co<strong>in</strong>cide when Ω collapses<br />

to a po<strong>in</strong>t.<br />

From the first two equation and the value of D 0,0 we get<br />

C 2,0 = ɛ 2 (n − 1)D 0,0 =<br />

D 2,1 = 2ɛD 0,0 = ɛ<br />

π n .<br />

ɛ(n − 1)<br />

4π n ,<br />

In order to f<strong>in</strong>d the value of C 4,1 and D 4,2 we consider the equations given by {c 3,0 =<br />

0, c 3,1 = 0, d 3,0 = 0, d 3,1 = 0}. The equation c 3,1 = 0 is the one given <strong>in</strong> (A.23), and the others<br />

are<br />

−ɛ(n − 2)C 2,0 − 24C 4,0 + (2n − 5)C 4,1 = 0,<br />

16C 4,0 − (n − 3)C 4,1 = 0,<br />

ɛ(n − 2)D 2,1 + C 4,1 − (n − 2)D 4,2 = 0.<br />

(Note that they co<strong>in</strong>cide with the ones <strong>in</strong> Theorem 4.3.5.) Solv<strong>in</strong>g the system given by these<br />

3 equations and (A.23), we get that it is compatible with solution<br />

C 4,0 =<br />

ɛ2<br />

32π n (n − 2)(n − 3), C 4,1 = ɛ2<br />

(n − 2),<br />

2πn D<br />

4,2 = 3ɛ2<br />

2π n .<br />

To f<strong>in</strong>d the value of the unknowns C 2c,q , D 2c,c with c ≥ 3, we have to solve the same<br />

equations as <strong>in</strong> the proof of Theorem 4.3.5. We can use the same relations if we first prove<br />

that C 4,0 , C 4,1 and D 4,2 also satisfies (A.12). We have to check it because variables C 4,0 , C 4,1


102 Appendix<br />

and D 4,2 here were obta<strong>in</strong>ed solv<strong>in</strong>g another l<strong>in</strong>ear system. But, it is straightforward verified.<br />

So, we get the same relation among the unknowns.<br />

Thus, from the equalities (A.13), (A.14), (A.15), (A.17) and (A.18), with r = n − 1, and<br />

the computation of C n,1 , C n,0 , D n,⌊n/2⌋ <strong>in</strong> the same way as <strong>in</strong> (A.19) we have<br />

χ(Ω) =<br />

n−1<br />

∑<br />

c=0<br />

c=0<br />

⎛<br />

ɛ c<br />

⎝<br />

π n<br />

∑c−1<br />

q=max{0,2c−n}<br />

1<br />

4 c−q ( c<br />

q<br />

)( n − c<br />

c − q<br />

)<br />

B ′ 2c,q(Ω) + c + 1<br />

Us<strong>in</strong>g the relation <strong>in</strong> (4.12) we get the stated expression<br />

⎛<br />

n−1<br />

∑ ɛ c ∑c−1<br />

( )( )<br />

χ(Ω) = ⎝<br />

1 c n − c<br />

π n 4 c−q B ′<br />

q c − q<br />

2c,q(Ω) + c + 1<br />

=<br />

=<br />

=<br />

n−1<br />

∑<br />

c=0<br />

+<br />

n−1<br />

∑<br />

c=0<br />

n−1<br />

∑<br />

c=0<br />

ɛ c (<br />

π n<br />

q=max{0,2c−n}<br />

∑c−1<br />

q=max{0,2c−n}<br />

2(c + 1)<br />

2<br />

ɛ c (<br />

π n<br />

2 Γ′ 2c,c(Ω)<br />

2 Γ′ 2c,c(Ω)<br />

⎞<br />

⎠+ ɛn (n + 1)!<br />

vol(Ω).<br />

⎞<br />

π n<br />

⎠ + ɛn (n + 1)!<br />

vol(Ω)<br />

1 c!(n − c)!q!(n − 2c + q)!(2c − 2q)!ω 2n−2c<br />

4 c−q B 2c,q (Ω)+<br />

q!(c − q)!(c − q)!(n − 2c + q)!<br />

c!(n − 2c + c)!(2c − 2c)!ω 2n−2c Γ 2c,c (Ω) ) + ɛn (n + 1)!<br />

vol(Ω)<br />

∑c−1<br />

q=max{0,2c−n}<br />

( )<br />

1 2c − 2q c!(n − c)!π<br />

n−c<br />

4 c−q B 2c,q (Ω)<br />

c − q (n − c)!<br />

π n−c<br />

+ (c + 1)!(n − c)!<br />

(n − c)! Γ 2c,c(Ω) ) + ɛn (n + 1)!<br />

π n vol(Ω)<br />

⎛<br />

ɛ c<br />

π c c! ∑c−1<br />

( )<br />

⎝<br />

1 2c − 2q<br />

4 c−q c − q<br />

q=max{0,2c−n}<br />

π n<br />

π n<br />

B 2c,q (Ω) + (c + 1)Γ 2c,c (Ω) ⎠ + ɛn (n + 1)!<br />

vol(Ω).<br />

⎞<br />

π n


Bibliography<br />

[AB08] S. Alesker and A. Bernig. The product on smooth and generalized valuations.<br />

arXiv:0904.1347v1, 2008.<br />

[Aba]<br />

J. Abardia. Average of mean curvature <strong>in</strong>tegral <strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong>. To appear<br />

<strong>in</strong> Advances <strong>in</strong> Geometry.<br />

[AF08] S. Alesker and J.H.G. Fu. Theory of valuations on manifolds. III: Multiplicative<br />

structure <strong>in</strong> the general case. Trans. Am. Math. Soc., 360(4):1951–1981, 2008.<br />

[AGS09] J. Abardia, E. Gallego, and G. Solanes. Gauss-bonnet theorem and crofton type formulas<br />

<strong>in</strong> <strong>complex</strong> <strong>space</strong> <strong>forms</strong>. Prepr<strong>in</strong>ts de la Universitat Autònoma de Barcelona,<br />

2009.<br />

[Ale01]<br />

[Ale03]<br />

[Ale04]<br />

S. Alesker. Description of translation <strong>in</strong>variant valuations on convex sets with solution<br />

of P. McMullen’s conjecture. Geom. Funct. Anal., 11(2):244–272, 2001.<br />

S. Alesker. Hard Lefschetz theorem for valuations, <strong>complex</strong> <strong>in</strong>tegral <strong>geometry</strong>, and<br />

unitarily <strong>in</strong>variant valuations. J. Differential Geom., 63(1):63–95, 2003.<br />

S. Alesker. SU (2)-<strong>in</strong>variant valuations. Milman, V. D. (ed.) et al., Geometric aspects<br />

of functional analysis. Papers from the Israel sem<strong>in</strong>ar (GAFA) 2002–2003. Berl<strong>in</strong>:<br />

Spr<strong>in</strong>ger. Lecture Notes <strong>in</strong> Mathematics 1850, 21-29 (2004)., 2004.<br />

[Ale06a] S. Alesker. Theory of valuations on manifolds. I: L<strong>in</strong>ear <strong>space</strong>s. Isr. J. Math.,<br />

156:311–339, 2006.<br />

[Ale06b] S. Alesker. Theory of valuations on manifolds. II. Adv. Math., 207(1):420–454, 2006.<br />

[Ale07a] S. Alesker. Theory of valuations on manifolds: a survey. Geom. Funct. Anal.,<br />

17(4):1321–1341, 2007.<br />

[Ale07b] S. Alesker. Theory of valuations on manifolds. IV: New properties of the multiplicative<br />

structure. Milman, Vitali D. (ed.) et al., Geometric aspects of functional<br />

analysis. Proceed<strong>in</strong>gs of the Israel sem<strong>in</strong>ar (GFA) 2004-2005. Berl<strong>in</strong>: Spr<strong>in</strong>ger. Lecture<br />

Notes <strong>in</strong> Mathematics 1910, 1-44 (2007)., 2007.<br />

[ÁPF04] J. C. Álvarez Paiva and E. Fernandes. What is a Crofton formula<br />

42:95–108, 2003/04.<br />

Math. Notae,<br />

[Ber08a] A. Bernig. A hadwiger-type theorem for the special unitary group. Geometric and<br />

Functional Analysis (to appear), 2008.<br />

[Ber08b] A. Bernig. <strong>Integral</strong> <strong>geometry</strong> under G 2 and Sp<strong>in</strong>(7). arXiv:0803.3885v1, 2008.<br />

[BF08] A. Bernig and J.H.G. Fu. Hermitian <strong>in</strong>tegral <strong>geometry</strong>. arXiv:0801.0711v5, 2008.<br />

103


104 Bibliography<br />

[Bla55]<br />

[Bla76]<br />

[Bor49]<br />

W. Blaschke. Vorlesungen über <strong>Integral</strong>geometrie. 3. Aufl. Berl<strong>in</strong>: VEB Deutscher<br />

Verlag der Wissenschaften 130 S., 44 Fig. im Text. , 1955.<br />

D.E. Blair. Contact manifolds <strong>in</strong> Riemannian <strong>geometry</strong>. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong>,<br />

1976. Lecture Notes <strong>in</strong> Mathematics, Vol. 509.<br />

A. Borel. Some remarks about Lie groups transitive on spheres and tori. Bull. Amer.<br />

Math. Soc., 55:580–587, 1949.<br />

[Bor50] A. Borel. Le plan projectif des octaves et les sphères comme e<strong>space</strong>s homogènes. C.<br />

R. Acad. Sci. Paris, 230:1378–1380, 1950.<br />

[GG00]<br />

[Gol99]<br />

C. Gorodski and N. Gusevskii. Complete m<strong>in</strong>imal hypersurfaces <strong>in</strong> <strong>complex</strong> hyperbolic<br />

<strong>space</strong>. Manuscripta Math., 103(2):221–240, 2000.<br />

W. M. Goldman. Complex hyperbolic <strong>geometry</strong>. Oxford Mathematical Monographs.<br />

The Clarendon Press Oxford University Press, New York, 1999. Oxford Science<br />

Publications.<br />

[Gra73] A. Gray. The volume of a small geodesic ball of a Riemannian manifold. Mich.<br />

Math. J., 20:329–344, 1973.<br />

[Gri78]<br />

P.A. Griffiths. Complex differential and <strong>in</strong>tegral <strong>geometry</strong> and curvature <strong>in</strong>tegrals<br />

associated to s<strong>in</strong>gularities of <strong>complex</strong> analytic varieties. Duke Math. J., 45:427–512,<br />

1978.<br />

[Had57] H. Hadwiger. Vorlesungen über Inhalt, Oberfläche und Isoperimetrie. Spr<strong>in</strong>ger-<br />

Verlag, Berl<strong>in</strong>, 1957.<br />

[Hsi98] P.H. Hsieh. L<strong>in</strong>ear submanifolds and bisectors <strong>in</strong> CH n . Forum Math., 10(4):413–434,<br />

1998.<br />

[KN69] S. Kobayashi and K. Nomizu. Foundations of differential <strong>geometry</strong>. Vol. II. Interscience<br />

Tracts <strong>in</strong> Pure and Applied Mathematics, No. 15 Vol. II. Interscience<br />

Publishers John Wiley & Sons, Inc., New York-London-Sydney, 1969.<br />

[KR97]<br />

[LS82]<br />

D.A. Kla<strong>in</strong> and G.-C. Rota. Introduction to geometric probability. Lezioni L<strong>in</strong>cee.<br />

Cambridge: Cambridge University Press. Rome: Accademia Nazionale dei L<strong>in</strong>cei,<br />

xiv, 178, 1997.<br />

R. Langev<strong>in</strong> and T. Shifr<strong>in</strong>. Polar varieties and <strong>in</strong>tegral <strong>geometry</strong>. Amer. J. Math.,<br />

104(3):553–605, 1982.<br />

[McM77] P. McMullen. Valuations and Euler-type relations on certa<strong>in</strong> classes of convex polytopes.<br />

Proc. London Math. Soc. (3), 35(1):113–135, 1977.<br />

[Mon85]<br />

[MS43]<br />

[MS83]<br />

[O’N83]<br />

S. Montiel. Real hypersurfaces of a <strong>complex</strong> hyperbolic <strong>space</strong>. J. Math. Soc. Japan,<br />

37(3):515–535, 1985.<br />

D. Montgomery and H. Samelson. Transformation groups of spheres. Ann. of Math.<br />

(2), 44:454–470, 1943.<br />

P. McMullen and R. Schneider. Valuations on convex bodies. In Convexity and its<br />

applications, pages 170–247. Birkhäuser, Basel, 1983.<br />

B. O’Neill. Semi-Riemannian <strong>geometry</strong>, volume 103 of Pure and Applied Mathematics.<br />

Academic Press Inc. [Harcourt Brace Jovanovich Publishers], New York, 1983.<br />

With applications to relativity.


Bibliography 105<br />

[Par02]<br />

H. Park. K<strong>in</strong>ematic formulas for the real sub<strong>space</strong>s of <strong>complex</strong> <strong>space</strong> <strong>forms</strong> of dimension<br />

2 and 3. PhD-thesis. University of Georgia, 2002.<br />

[Rum94] M. Rum<strong>in</strong>. Formes différentielles sur les variétés de contact. J. Differential Geom.,<br />

39(2):281–330, 1994.<br />

[San52] L. A. Santaló. <strong>Integral</strong> <strong>geometry</strong> <strong>in</strong> Hermitian <strong>space</strong>s. Amer. J. Math., 74:423–434,<br />

1952.<br />

[San04]<br />

[Sch93]<br />

[Sol06]<br />

[War71]<br />

L. A. Santaló. <strong>Integral</strong> <strong>geometry</strong> and geometric probability. Cambridge Mathematical<br />

Library. Cambridge University Press, Cambridge, second edition, 2004. With a<br />

foreword by Mark Kac.<br />

R. Schneider. Convex bodies: the Brunn-M<strong>in</strong>kowski theory, volume 44 of Encyclopedia<br />

of Mathematics and its Applications. Cambridge University Press, Cambridge,<br />

1993.<br />

G. Solanes. <strong>Integral</strong> <strong>geometry</strong> and the Gauss-Bonnet theorem <strong>in</strong> constant curvature<br />

<strong>space</strong>s. Trans. Amer. Math. Soc., 358(3):1105–1115 (electronic), 2006.<br />

F.W. Warner. Foundations of differentiable manifolds and Lie groups. Scott, Foresman<br />

and Co., Glenview, Ill.-London, 1971.


Notation<br />

( , ) : Hermitian product <strong>in</strong> C n+1 def<strong>in</strong>ed at (1.2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

( , ) ɛ : Hermitian product <strong>in</strong> CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

〈 , 〉 ɛ : Hermitian metric of CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

α i : real part of a dual form of a J-mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

α ij : real part of a connection form of a J-mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . . . . . . . 16<br />

B k,q (Ω) : Hermitian <strong>in</strong>tr<strong>in</strong>sic volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

β i : imag<strong>in</strong>ary part of a dual form of a J-mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . . . . . . . . 16<br />

β ij : imag<strong>in</strong>ary part of a connection form of a J-mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . 16<br />

C n : standard Hermitian <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

CH n : <strong>complex</strong> hyperbolic <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

CK n (ɛ) : <strong>complex</strong> <strong>space</strong> form with constant holomorphic curvature 4ɛ . . . . . . . . . . . . . . . . . 8<br />

CP n : <strong>complex</strong> projective <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

cos ɛ (α) : generalized cos<strong>in</strong>e function with angle α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

cot ɛ (α) : generalized cotangent function with angle α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

D :<br />

distribution def<strong>in</strong>ed by the normal vector and the <strong>complex</strong> structure <strong>in</strong> a hypersurface<br />

<strong>in</strong> a Kähler manifold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

dL r : density of the <strong>space</strong> of <strong>complex</strong> planes with <strong>complex</strong> dimension r . . . . . . . . . . . . . . . 20<br />

E : bisector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

Γ k,q (Ω) : Hermitian <strong>in</strong>tr<strong>in</strong>sic volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

G C n,r : Grassmannian of <strong>complex</strong> r-planes <strong>in</strong> C n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

G n,k,p (C n ) : Grassmannian of (k, p)-planes <strong>in</strong> C n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

H n : real hyperbolic <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

H : sub<strong>space</strong> of C n+1 which def<strong>in</strong>es the po<strong>in</strong>ts <strong>in</strong> CK n (ɛ) <strong>in</strong> the projective model . . . . . . 9<br />

J : <strong>complex</strong> structure of a <strong>complex</strong> manifold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

K(V ) : convex compact doma<strong>in</strong>s <strong>in</strong> the vector <strong>space</strong> V . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

107


108 Notation<br />

L R r : <strong>space</strong> of totally real planes with dimension r <strong>in</strong> CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

L R r : totally real plane <strong>in</strong> L R r . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

L C r : <strong>space</strong> of <strong>complex</strong> planes with <strong>complex</strong> dimension r <strong>in</strong> CK n (ɛ) . . . . . . . . . . . . . . . . . . . 19<br />

L C r : <strong>complex</strong> plane <strong>in</strong> L C r . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

L (r)<br />

q : <strong>space</strong> of <strong>complex</strong> q-planes conta<strong>in</strong>ed <strong>in</strong> a fixed <strong>complex</strong> r-plane . . . . . . . . . . . . . . . . . 25<br />

: <strong>space</strong> of <strong>complex</strong> r-planes conta<strong>in</strong><strong>in</strong>g a fixed <strong>complex</strong> q-plane . . . . . . . . . . . . . . . . . . 25<br />

L C r[q]<br />

M n×n (C) : square matrices of order n with <strong>complex</strong> valued entries . . . . . . . . . . . . . . . . . . . 11<br />

M i (S) : i-th mean curvature <strong>in</strong>tegral of a hypersurface S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

M D i<br />

(S) : i-th mean curvature <strong>in</strong>tegral restricted to the distribution D . . . . . . . . . . . . . . . . 40<br />

µ k,q : Hermitian <strong>in</strong>tr<strong>in</strong>sic volumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

N(Ω) : normal fiber bundle of a doma<strong>in</strong> Ω ⊂ CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

O n : area of the sphere of radius 1 <strong>in</strong> the standard Euclidean <strong>space</strong> of dimension n . . . 26<br />

ϕ i : dual form of a J-mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

ϕ ij : connection form of a J-mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

P U(n) : projective unitary group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

RH n : real projective <strong>space</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

S(CK n (ɛ)) : unit tangent bundle of CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

σ i (A) : i-th symmetric elementary function of the bil<strong>in</strong>eal form A . . . . . . . . . . . . . . . . . . . . 30<br />

s<strong>in</strong> ɛ (α) : generalized s<strong>in</strong>e function with angle α . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

U k,p : valuation def<strong>in</strong>ed on C n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

U(p, q) : unitary group of <strong>in</strong>dex p, q . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

U ɛ (n) : isometry group of CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

V i (Ω) : i-th <strong>in</strong>tr<strong>in</strong>sic volume of the doma<strong>in</strong> Ω ⊂ R n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

Val(V ) : cont<strong>in</strong>uous translation <strong>in</strong>variant valuations on the vector <strong>space</strong> V . . . . . . . . . . . 31<br />

Val k (V ) : homogeneous valuations of degree k <strong>in</strong> Val(V ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

Val + (V ) : even valuations <strong>in</strong> Val(V ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

Val − (V ) : odd valuations <strong>in</strong> Val(V ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

Val G (V ) : <strong>in</strong>variant valuations under the action of the group G <strong>in</strong> Val(V ) . . . . . . . . . . . . 31<br />

Val U(n) (C n ) : cont<strong>in</strong>uous translation and U(n)-<strong>in</strong>variant valuations on C n . . . . . . . . . . . . 34<br />

ω n : volume of the ball of radius 1 <strong>in</strong> the standard Euclidean <strong>space</strong> of dimension n . . . 38<br />

ω i0 : dual form of an orthonormal mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

ω ij : connection form of an orthonormal mov<strong>in</strong>g frame on CK n (ɛ) . . . . . . . . . . . . . . . . . . . . 16


Index<br />

2-po<strong>in</strong>t homogeneous <strong>space</strong>, 14<br />

3-po<strong>in</strong>t homogeneous <strong>space</strong>, 14<br />

aff<strong>in</strong>e surface area, 30<br />

Alesker, 29, 32, 34, 36, 37, 58<br />

almost <strong>complex</strong> structure, 7<br />

angle between two planes, 9<br />

Bergmann metric, 10<br />

Bernig, 36–38, 61<br />

bisector, 18, 82, 83<br />

Blaschke, 29<br />

cohomogenity 1 hypersurface, 83<br />

coisotropic plane, 88<br />

<strong>complex</strong><br />

hyperbolic <strong>space</strong>, 8<br />

manifold, 7<br />

projective <strong>space</strong>, 7, 8<br />

<strong>space</strong> form, 7, 8, 14, 18, 36<br />

sp<strong>in</strong>e of a bisector, 82<br />

structure, 7<br />

submanifold, 17<br />

sub<strong>space</strong>, 17, 81<br />

cont<strong>in</strong>uous valuation, 30<br />

Crofton formula, 32<br />

even valuation, 31<br />

Fu, 36–38, 61<br />

Fub<strong>in</strong>i-Study metric, 10<br />

generic position, 48<br />

geodesic ball, 18<br />

Hadwiger, 32<br />

Hadwiger Theorem, 32, 36, 37<br />

Hausdorff<br />

distance, 30<br />

metric, 30<br />

Hermitian<br />

<strong>in</strong>tr<strong>in</strong>sic volume, 40, 61<br />

product, 7<br />

holomorphic<br />

angle, 9<br />

atlas, 7<br />

curvature, 8<br />

section, 8<br />

homogeneous<br />

hypersurface, 83<br />

<strong>space</strong>, 13, 19<br />

valuation of degree k, 31<br />

horizontal lift, 10<br />

<strong>in</strong>tr<strong>in</strong>sic volume, 29–31<br />

<strong>in</strong>variant<br />

density, 19, 20<br />

with respect to a group, 31<br />

isometry group, 11, 14, 34<br />

isotropy group of a <strong>complex</strong> r-plane, 19<br />

J-bases, 14<br />

J-mov<strong>in</strong>g frame, 14<br />

(k, p)-plane, 81<br />

Kähler<br />

form, 8<br />

manifold, 8<br />

k<strong>in</strong>ematic formula, 33<br />

Lagrangian plane, 36, 90<br />

l<strong>in</strong>ear<br />

(k, p)-plane, 81<br />

submanifold, 81<br />

McMullen, 32<br />

Theorem, 31<br />

mean curvature <strong>in</strong>tegrals, 30, 31, 33<br />

monotone valuation, 31<br />

mov<strong>in</strong>g frame, 14<br />

odd valuation, 31<br />

Park, 38–40, 75<br />

Quermas<strong>in</strong>tegrale, 26<br />

109


110 Index<br />

real<br />

hyperbolic <strong>space</strong>, 8<br />

projective <strong>space</strong>, 8<br />

<strong>space</strong> form, 8, 18<br />

real sp<strong>in</strong>e of a bisector, 82<br />

regular doma<strong>in</strong>, 16<br />

reproductive<br />

formula, 33<br />

property, 45, 56<br />

restricted mean curvature <strong>in</strong>tegral, 40<br />

r-pla <strong>complex</strong>, 17<br />

r-pla totalment real, 17<br />

Rum<strong>in</strong><br />

derivative, 61<br />

operator, 61<br />

second fundamental form, 30<br />

slide of a bisector, 82<br />

smooth valuation, 31<br />

on a manifold, 37<br />

<strong>space</strong> constant<br />

holomorphic curvature, 7, 8, 14, 18<br />

sectional curvature, 8, 18<br />

sp<strong>in</strong>al surface, 82<br />

standard Hermitian <strong>space</strong>, 8<br />

Ste<strong>in</strong>er formula, 29, 30<br />

symmetric elementary function, 30, 40<br />

symplectic form, 38<br />

template method, 33<br />

totally<br />

geodesic<br />

submanifold, 17<br />

sub<strong>space</strong>, 81<br />

real<br />

plane, 9<br />

submanifold, 17<br />

sub<strong>space</strong>, 17, 81<br />

umbilical hypersurface, 18<br />

translation <strong>in</strong>variant valuation, 31<br />

trigonometric generalized functions, 10<br />

unimodular group, 13<br />

unit normal bundle, 16, 17<br />

unit tangent bundle, 15<br />

U(p, q), 11<br />

valuation, 29

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!